4.5.5, 1 I v Ema”; , u:- .1... Suit! 7 . 5.105! 335.55“ .0 fungi?» 3... c: .wnum-i. 4 ulL—t. if . . . “flan; 74‘s.. ..!£.&i.&..th".!§|l§ it... fin. I 1.1.. 5...}: ’.{I33 . Re 25 a! .2. {£15539 I!" 5).... Kiel-3. 3a 4.): Juxlfli . 1! .. - tugi...!1i 1:...2123. it. A¢§....it.€ «Aiken! a 2“ 3 tie‘ ‘ (2:5 1.19:5... \ \ 1-i1;’r.vidni;§:rfl§ 1.13. :2 a... zi€a .. 319.3... .129}- .. 3 l IBRARY 83:0 Michigan State UnWemmy This is to certify that the dissertation entitled ELECTRICALLY-ACTIVE POLYANILINE COATED MAGNETIC NANOPARTICLES: A NOVEL CONCENTRATOR AND NANOSTRUCTURED TRANSDUCER IN BIOSENSOR DEVICES presented by SUDESHNA PAL has been accepted towards fulfillment of the requirements for the PhD. degree in Biosystems Engineering Major Professor% Signature Q/J4/o? Date MSU is an Affirmative Action/Equal Opportunity Employer PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 K:IProlecc&PrelelRC/DateDm.indd ELECTRICALLY-ACTIVE POLYANILINE COATED MAGNETIC NANOPARTICLES: A NOVEL CONCENTRATOR AND NANOSTRUCTURED TRANSDUCER IN BIOSENSOR DEVICES By Sudeshna Pal A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Biosystems Engineering 2009 ABSTRACT ELECTRICALLY-ACTIVE POLYANILINE COATED MAGNETIC NANOPARTICES: A NOVEL CONCENTRATOR AND NANOSTRUCTURED TRANSDUCER IN BIOSENSOR DEVICES By Sudeshna Pal Nanomaterial based biosensors are emerging as sensitive, Specific, efficient and rapid diagnostic tools for the detection of pathogenic microorganisms in biodefense and diagnostics. This research demonstrates the novel application of a nanostructured material in the dual function of a magnetic concentrator and a transducer in biosensor devices. The nanostructured materials are synthesized from aniline monomer made electrically active by acid doping coating the surface of gamma iron oxide cores. The synthesized electrically active polyaniline coated magnetic (EAPM) nanoparticles are adapted in antibody based and DNA based biosensors for the detection of Bacillus anthracis as a model pathogen. The antibody based detection involves immunomagnetic concentration of targets using biologically modified EAPM nanoparticles followed by their direct application to a charge transfer biosensor. Experimental results indicate that the biosensor is able to detect B. anthracis spores at concentrations as low as 4.2 X 102 spores per ml in 16 min. The DNA based detection mechanism encompasses sandwiched hybridization of DNA targets onto probe labeled EAPM nanoparticles which is succeeded by electrochemical detection of the EAPM DNA hybrids on screen printed carbon electrodes. The sensitivity of the DNA biosensor is assessed to be 0.01ng/ul of PCR amplified B. anthracis pag A gene fragment in a total detection time of 60 min. To my parents and my husband, Tuhin iii ACKNOWLEDGMENTS I sincerely thank everyone who contributed to the completion of this research. I express my deepest gratitude to Dr. Evangelyn Alocilja for being more than an advisor to me. I am grateful for her guidance, time, encouragement and appreciation for even the smallest achievements, and her prayers during my graduate program. I am thankful to the members of my Ph.D. guidance committee: Dr. David Tomanek, Dr. Shantanu Chakrabartty, Dr. Daniel Guyer and Dr. Frances Downes for their suggestion and support. I thank all past and present members of the Biosensors research group (Deng, Edith, Emma, Finny, Maria, Lisa, Michelle, Michael, Shannon, Srini, Yang, Yilun and Zarini) for sharing a pleasant work place. A special thank you to Ms. Edith Torres Chavolla for her guidance in molecular biology and for the help she extended to me. I am also thankful to Ms. Hanna Miller and all the other undergraduate assistants who have worked with me over the past few years. I owe my gratitude to my parents and my sisters for the unconditional love they have showered upon me throughout the years. Their unceasing support and concern for me will always be my strength. Last but not the least; I would like to express my heartfelt gratitude to my husband for being my friend, philosopher, and guide in our wonderful journey together. Thank you very much for your love and support!! Sudeshna Pal iv TABLE OF CONTENTS List of Tables ix List of Figures xii Chapter 1: INTRODUCTION 1 Chapter 2: LITERATURE REVIEW 4 2.1 Bacillus anthracis and Anthrax ........................................................................... 4 2.1.1 Virulence of Anthrax .............................................................................. 4 2.1.2 Routes of Infection .................................................................................. 6 2.1.3 Anthrax and Biosecurity ......................................................................... 7 2.1.4 Traditional Identification Methods ......................................................... 8 2. 2 Biosensors in Pathogen Detection ..................................................................... 10 2.2.1 Introduction ........................................................................................... 10 2.2.2 Biosensor Transducing Mechanisms .................................................... 12 2.2.2.1 Mechanical biosensors .................................................................... 12 2.2.2.2 Optical biosensors ........................................................................... 17 2.2.2.3 Electrochemical biosensors ............................................................ 21 2.2.2.4 Magnetic biosensors ....................................................................... 25 2.2.3 Biological Sensing Elements ................................................................ 27 2. 3 NanoBiosensors and Nanostructured Transducers ........................................... 30 2.3.1 Nanowire Based Biosensors ................................................................. 30 2.3.1.1 Silicon nanowires ............................................................................ 31 2.3.1.2 Conducting poymer nanowires ....................................................... 32 2.3.2 Nanotube Based Biosensors .................................................................. 34 2.3.2.1 Carbon nanotubes ........................................................................... 34 2.3.2.2 Conducting poymer nanotubes ....................................................... 36 2.3.3 NanOparticle Based Biosensors ............................................................. 37 2.3.3.1 Magnetic nanoparticles ................................................................... 37 2.3.3.2 Gold nanoparticles .......................................................................... 39 2.3.3.3 Semiconductor nanoparticles .......................................................... 42 2.3.3.4 Silica nanoparticles ......................................................................... 44 2.3.4 Nanoporous Material Based Biosensors ............................................... 46 2.3.5 Nanorod and Nanosphere Based Biosensors ........................................ 47 2. 4 Polyaniline — A Conducting Polymer ................................................................ 49 2.4.1 Introduction ........................................................................................... 49 2.4.2 Structural and Electrical Properties ...................................................... 51 2.4.2.1 Oxidation states .............................................................................. 52 2.4.2.2 Electronic conduction ..................................................................... 54 2.4.3 Electrochemistry and Redox Switching ................................................ 58 2.4.3.1 Cyclic voltammetry ........................................................................ 61 2.4.4 Electrically-Active Magnetic Polyaniline ............................................. 63 2.4.5 Polyaniline in Biosensor Applications .................................................. 67 Chapter 3: RESEARCH HIGHLIGHTS 70 3.1 Research Novelty ............................................................................................... 70 3.2 Research Significance ........................................................................................ 72 3.3 Hypothesis .......................................................................................................... 73 3.4 Research Objectives ........................................................................................... 73 Chapter 4: RESEARCH MATERIALS AND METHODS - - 74 4.] Objective 1 ......................................................................................................... 74 4.1.1 EAPM Nanoparticle Synthesis ............................................................. 74 4.1.2 EAPM Nanoparticle Characterization .................................................. 75 4.1.2.1 Magnetic characterization ............................................................... 75 4.1.2.2 Electrical characterization .............................................................. 75 4.1.2.3 Structural characterization .............................................................. 76 4.1.2.4 Spectral analysis ............................................................................. 76 4.2 Objective 2 ......................................................................................................... 77 4.2.1 Biosensor Design and Data Collection ................................................. 77 4.2.2 EAPM Based Immunosensor Fabrication ............................................. 79 4.2.2.1 Chemicals and reagents .................................................................. 79 4.2.2.2 EAPM antibody modification ......................................................... 79 4.2.2.3 Biosensor capture pad functionalization ......................................... 80 4.2.2.4 Application and absorption pad preparation ................................... 81 4.2.2.5 Sensor assembly .............................................................................. 81 4.2.2.6 Antibody attachment confirmation studies ..................................... 82 4.3 Objective 3 ......................................................................................................... 82 4.3.1 Bacterial Culture and Plating ................................................................ 83 4.3.1.1 B. anthracis sporulation protocol ................................................... 83 4.3.2 Immunomagnetic Concentration Using EAPM NPs ............................ 84 4.3.3 Biosensor Testing .................................................................................. 85 4.3.4 Biosensor Sensitivity Study .................................................................. 85 4.3.5 Biosensor Specificty Analysis .............................................................. 86 4.3.6 Biosensor Testing in Food Matrices ..................................................... 86 4.3.7 Statistical Analysis ................................................................................ 87 4.3.8 Confirmation and EAPM Capture Efficiency Calculations .................. 87 4. 4 Objective 4 ......................................................................................................... 88 4.4.1 Biosensor Design and Data Collection ................................................. 88 4.4.2 Chemicals and Reagents ....................................................................... 89 4.4.3 Selection and Analysis of DNA Primers and Probes ............................ 90 4.4.4 Extraction, Amplification and Characterization of B. anthracis DNA. 90 4.4.4.1 DNA extraction ............................................................................... 90 4.4.4.2 Polymerase chain reaction (PCR) amplification and Optimization. 92 4.4.4.3 PCR product purification ................................................................ 92 Vi 4.4.4.4 PCR product characterization ......................................................... 93 4.4.5 EAPM-DNA Probe Labeling Study ...................................................... 93 4.4.5.1 EAPM-DNA modification .............................................................. 93 4.4.5.2 Confirmation and optimization ....................................................... 94 4.4.6 SPCE Surface Modification .................................................................. 95 4.5 Objective 5 ......................................................................................................... 95 4.5.1 Sandwiched Hybridization of DNA Targets on EAPM NPs ................ 95 4.5.1.1 Sandwiched hybridization protocol ................................................ 95 4.5.1.2 Optimization of hybridization conditions and confirmation .......... 96 4.5.2 Electrochemical Detection Using SPCE Biosensor .............................. 97 4.5.2.1 Electrochemical Characterization of EAPM NPS ........................... 97 4.5.2.2 Detection of EAPM-Target-Biotin Hybrids ................................... 97 4.5.3 SPCE Biosensor Sensitivity Study ........................................................ 98 4.5.4 Statistical Analysis ................................................................................ 98 4.5.5 Specificity Analysis .............................................................................. 99 Chapter 5: RESULTS AND DISCUSSION 100 5.1 Objective 1 ....................................................................................................... 100 5.1.1 EAPM Nanoparticle Characterization ................................................ 100 5.2 Objective 2 ....................................................................................................... 112 5.2.1 EAPM Based Immunosensor Fabrication ........................................... 112 5.2.1.1 Confirming EAPM antibody modification ................................... 112 5.2.1.2 Confirming biosensor capture pad functionalization .................... 113 5.2.2 EAPM Based Immunosensor Detection Concept ............................... 116 5.2.2.1 Validation of detection concept .................................................... 118 5.3 Objective 3 ....................................................................................................... 122 5.3.1 Immunosensor Performances Using Different EAPM Types ............. 122 5.3.2 Immunosensor Sensitivity Study ........................................................ 125 5.3.3 Immunosensor Specificity Analysis ................................................... 128 5.3.4 Immunosensor Testing in Food Matrices ........................................... 131 5.3.5 Confirmation and EAPM Capture Efficiency Determination ............. 134 5.4 Objective 4 ....................................................................................................... 138 5.4.] Selection and Analysis of DNA Primers and Probes .......................... 138 5.4.2 Extraction, Amplification and Characterization of B. anthracis DNA138 5.4.3 Confirming EAPM-DNA Probe Modification .................................... 141 5.4.4 EAPM Based Electrochemical DNA Biosensor Detection Concept .. 143 5.5 Objective 5 ....................................................................................................... 145 5.5.1 Sandwiched Hybridization of DNA Targets on EAPM NPS ............. 145 5.5.1.1 Determination of hybridization conditions ................................... 145 5.2.1.2 Confirming sandwiched hybridization on EAPM NPS ................. 147 5.5.2 Electrochemical Detection Using SPCE Biosensor ............................ 149 5.5.2.1 Electrochemical characterization of EAPM NPS .......................... 149 5.5.2.2 Electrochemical detection of EAPM-Target-Biotin DNA hybrid 153 5.5.3 SPCE Biosensor Sensitivity Study ...................................................... 155 5.5.4 Specificity Analysis ............................................................................ 160 vii Chapter 6: CONCLUSION AND FUTURE RESEARCH - -_ 162 Appendix A: STATISTICAL ANALYSIS RESULTS 165 A. I Anova Analysis of Different EAPM Based Immunosensors ............................. 165 A. 2 Anova Analyszs for Immunosensor Sensitivity ................................................. 168 A3 Anova Analysis for Immunosensor Specificity ................................................. I 70 A4 Anova Analysis for Lettuce Testing .................................................................. 1 71 A5 Anova Analysis for Ground Beef Testing ......................................................... 1 73 A6 Anova Analysis for Whole Milk Testing ........................................................... 175 A. 7 Anova Analysis for SPCE Biosensor Sensitivity .............................................. 1 77 Appendix B: DATA - 179 REFERENCES 183 viii LIST OF TABLES Table 2-1. Chemical structure and conductivity of typical conducting polymers (Adapted and modified from Rahman et al., 2008). ....................................................... 50 Table 2-2. Review of electrically-active magnetic polyaniline nanostructures. ............... 66 Table 3-1. Novelty of research as compared to existing literature. .................................. 71 Table 4-1. Dimensions of the EAPM based immunosensor. ............................................ 79 Table 4-2. Sequence information of B. anthracis primer pairs and probes. ..................... 91 Table 5-1. Magnetic parameters measured for the EAPM NPs ...................................... 102 Table 5-2. Conductivity values measured for the EAPM NPS. ...................................... 104 Table 5-3. Capture efficiency (CE) of immno-EAPMS in pure spore suspensions of B. anthracis.estimated from plate count data .................................................... 136 Table 5-4. Capture efficiency (CE) of immuno-EAPMS in romaine lettuce, whole milk, and ground beef contaminated with B. anthracis spores estimated from plate count data ...................................................................................................... 137 Table 5-5. Optimum concentration of PCR components. ............................................... 139 Table 6-1. Specifications of EAPM based biosensors. ................................................... 164 Table A-1. Type 3 tests of fixed effects (Anova analysis of different EAPM based immunosensors) ........................................................................................... 165 Table A-2. Tests of effect slices (Anova analysis of different EAPM based immunosensors) ........................................................................................... 165 Table A-3. Estimates (Anova analysis of different EAPM based immunosensors) ....... 166 Table A-4. Least squares means (Anova analysis of different EAPM based immunosensors) ........................................................................................... 1 66 ix Table A-5. Differences of least squares means (Anova analysis of different EAPM based immunosensors) ........................................................................................... l 67 Table A-6. Type 3 tests of fixed effects (Anova analysis for immunosensor sensitivity) ..................................................................................................................... 168 Table A-7. Estimates (Anova analysis for immunosensor sensitivity) ........................... 168 Table A-8. Least squares means (Anova analysis for immunosensor sensitivity) ......... 168 Table A-9. Differences of least squares means (Anova analysis for immunosensor sensitivity) ................................................................................................... 1 69 Table A-10. Table A-1 1. Table A-12. Table A-13. Table A-l4. Table A-15. Table A—16. Table A-17. Table A-18. Table A-19. Table A-20. Table A-21. Type 3 tests of fixed effects (Anova analysis for immunosensor specificity) ..................................................................................................................... 170 Estimates (Anova analysis for immunosensor specificity) ......................... 170 Type 3 tests of fixed effects (Anova analysis for lettuce testing) .............. 171 Least squares means (Anova analysis for lettuce testing) .......................... 171 Differences of least squares means (Anova analysis for lettuce testing) 172 Type 3 tests of fixed effects (Anova analysis for ground beef testing) ...... 173 Least squares means (Anova analysis for ground beef testing) .................. 173 Differences of least squares means (Anova analysis for ground beef testing) ..................................................................................................................... 174 Type 3 tests of fixed effects (Anova analysis for whole milk testing) ....... 175 Least squares means (Anova analysis for whole milk testing) ................... 175 Differences of least squares means (Anova analysis for whole milk testing) ..................................................................................................................... 176 Type 3 tests of fixed effects (Anova analysis for SPCE biosensor sensitivity) ..................................................................................................................... 177 Table A-22. Estimates (Anova analysis for SPCE biosensor sensitivity) ...................... 177 Table A-23. Least squares means (Anova analysis for SPCE biosensor sensitivity) ..... 177 Table A-24. Differences of least squares means (Anova analysis for SPCE biosensor sensitivity) ................................................................................................... l 78 xi LIST OF FIGURES Figure 2-1. Plasmids and genes associated with Virulence of B. anthracis (adapted from Mock and F ouet, 2001). ................................................................................ 5 Figure 2-2. Mode of action of B. anthracis toxin proteins ................................................. 6 Figure 2-3. Schematic representation of a biosensor. ....................................................... 10 Figure 2-4. Recent trends in pathogen detection (adapted from Lazcka et al. , 2007). ..... 1 1 Figure 2-5. Schematic of the polyaniline nanowire biosensor for B. cereus detection (Pal et al., 2008). ................................................................................................... 33 Figure 2-6. TEM images of (A) single-walled carbon nanotubes and (B) multi-walled carbon nanotubes. .......................................................................................... 35 Figure 2-7. Schematic representation of Au nanoparticle based bio barcode assay (Zhang et al., 2009) .................................................................................................... 42 Figure 2-8. Structure of polyaniline, x = degree of polymerization, [y + (1 -y)] =1] (Ray et al., 1989) ........................................................................................................ 53 Figure 2-9. Structure of emeraldine (polyaniline), (a) before protonation and (b)-(d) after protonation, (b) formation of bipolarons, (0) formation of polarons, and (d) the separate polarons which result in a polaron lattice (adapted from Stafstrom et al., 1987) .................................................................................... 56 Figure 2-10. Band structure of conjugated polymers showing polaronic and bipolaronic states ............................................................................................................... 57 Figure 2-11. CVs of polyaniline films recorded at pH values of (a) 1.0, (b) 2.0, (c) 2.5, (d) 3.0 and (e) 4.0 (Prakash, 2001). ..................................................................... 60 Figure 2-12. Typical cyclic voltammogram (I-E curve) for a redox system .................... 63 Figure 2-13. Synthesis of electrically-active magnetic polyaniline (Li et al., 2007). ....... 64 Figure 2-14. Polymer ligand interactions to iron oxide particle sufaces (L = Ligand) (Kryszewski and Jeszka, 1998) ...................................................................... 65 xii Figure 4-1. Schematic representation of polyaniline coating of gamma-iron oxide nanoparticles .................................................................................................. 75 Figure 4-2. Schematic of (A) EAPM based immunosensor architecture and (B) data collection system for detection of B. anthracis spores. ................................. 78 Figure 4-3. Modification of EAPM NPS with monoclonal antibodies. ............................ 80 Figure 4-4. Functionalization of biosensor capture pad for attachment of polyclonal antibodies. ...................................................................................................... 81 Figure 4-5. Screen printed carbon electrode biosensor. .................................................... 89 Figure 4-6.Crosslinking chemistry between the detector DNA probe and magnetic EAPM NPS ................................................................................................................. 94 Figure 5-1. Experimental M-H curves of the four different EAPM NPS and unmodified Fe203 NPS at 300K. ..................................................................................... 101 Figure 5-2. Experimental F C-ZFC magnetization curves for 1:0.6 EAPM NPS at 100 Oe ...................................................................................................................... 102 Figure 5-3. Room temperature electrical conductivity of the EAPM NPS ..................... 104 Figure 5-4. TEM and electron diffraction images of y-Fe203 NPS. ............................... 105 Figure 5-5. TEM and electron diffraction images of EAPM NPS (left) 1: 0.1 EAPM; (right) 1:0.4 EAPM .................................................................................... 106 Figure 5-6. TEM and electron diffraction images of EAPM NPs (left) 1: 0.6 EAPM; (right) 1:0.8 EAPM .................................................................................... 107 Figure 5-7. Elemental analysis of the 1:0.6 EAPM NPS using energy dispersive spectroscopy (EDS) in SEM. ..................................................................... 109 Figure 5-8. UV-VIS spectra absorption spectra of pure polyaniline and EAPM NPs.... 111 Figure 5-9. UV spectrum of pure anti-B. anthracis IgG molecules and unreacted IgG molecules after magnetic separation from EAPM NPS ............................. 113 xiii Figure 5-10. Figure 5-11. Figure 5-12. Figure 5-13. Figure 5-14. Figure 5-15. Figure 5-16. Figure 5-17. Figure 5-18. Figure 5-19. Figure 5-20. Figure 5-21. Laser scanning confocal microssope images of antibody modified biosensor capture pad with FITC anti-goat IgG label (left) and capture pad without FITC anti—goat IgG label (right) (scale bar = 20pm) ................................. 115 Schematic representation of the EAPM based immunosensor for detection of B. anthracis spores ..................................................................................... 117 Amperometric response of the EAPM based immunosensor in the presence and absence of the target antigen. .............................................................. 120 Resistance changes in the EAPM immunosensor in the presence and absence of the target antigen calculated from the measured amperometric response ................................................................................................................... 121 Comparison of biosensor resistance response of the four EAPM NP at three different spore concentrations (Mean resistance d: SD, n = 3) . ................ 124 EAPM based immunosensor resistance response in pure spore suspensions of B. anthracis (mean resistance 3: SD, n = 3). Different superscripts over the bars indicate significantly different resistance response (P < 0.05) 127 Comparison of immunosensor resistance response in pure spore suspensions of B. anthracis and pure cultures of generic E. coli and S. Enteritidis. Different superscripts over bars indicate sgnificantly different resistance response (P < 0.05) .................................................................................... 130 Resistance response of the EAPM based immunosensor in food matrices contaminated with B. anthracis spores. For each food sample, different superscripts over bars indicate significantly different resistance response(P < 0.05). ....................................................................................................... 133 Gel electrophoresis of PCR amplified Bacillus anthracis. (M- Marker, BAl and BA2- B. anthracis, NC- Negative control) ......................................... 140 Fluorescence signal of pure Ph-PRO probes and unreacted Ph-PRO probes after magnetic separation from the EAPM NPS ......................................... 142 Schematic representation of the EAPM based electrochemical DNA biosensor for detecting B. anthracis pag gene ........................................... 144 Normalized fluorescence signal of PCR DNA targets before and after hybridization at different hybridization temperatures. .............................. 146 xiv Figure 5-22. Normalized fluorescence signal of PCR DNA targets before and after hybridization at different hybridization times. .......................................... 147 Figure 5-23. Change in fluorescence of PRO-Bio probes in the presence and absence of PCR targets during sandwiched hybridization ........................................... 148 Figure 5-24. Cyclic voltammograms of (a) EAPM NPS on SPCE and (b) bare SPCE in 0.1 MHCl at 20 mV/s ................................................................................ 151 Figure 5-25. Cyclic voltammograms of EAPM NPs in 0.1 M HCl at scan rates of 20, 50, 100, 150 and 200 mV/s. Inset- Plots of anodic and cathodic peak current vs. scan rate. .................................................................................................... 152 Figure 5-26. Electrochemical responses of EAPM captured targets DNA hybrids on SPCE, of bare SPCE, and of streptavidin modified SPCE in 0.1 M HCl at 20 mV/s ............................................................................................................ 154 Figure 5-27. Cyclic voltammograms of the EAPM based SPCE biosensor in different concentrations of PCR target DNA in 0.1M HCl at a scan rate of 20 mV/s. ..................................................................................................................... 156 Figure 5-28. Anodic (oxidation) peaks of PCR target concentrations ranging from 1ng/ul to 0 ng/ul, inset: anodic peak of 10 ng/ul PCR target (obtained from cyclic voltammograms) .......................................................................................... 1 57 Figure 5-29. CV mediated anodic peak current at different PCR target concentrations from three experimental trials (mean current i SD, n = 3). ........................ 158 Figure 5-30. Change in fluorescence intensity of non-complementary sequences before and after sandwiched hybridization on EAPM NPS. ................................... 161 Figure 8-1. SEM images of (A) Iron oxide NPs and (B) EAPM NPS ............................ 179 Figure B-2. Immunosensor responses of different EAPM concentrations in B. cereus. 180 Figure B-3. Immunosensor responses of different B. cereus antibody concentrations .. 181 Figure B-4. Light microscopy images of B. anthracis spores at after two different incubation periods (Final spore count — 4.2 X 108 spores/ml) ..................... 182 XV CHAPTER 1: INTRODUCTION In 2001, the anthrax cases associated with the intentional distribution of Bacillus anthracis spores through the postal system highlighted the danger posed to global citizens from bioterrorism related attacks. A total of 22 cases of anthrax were identified by the Centers for Disease Control and Prevention (CDC), of which 11 were confirmed to be inhalational and the remaining were determined to be cutaneous (CDC, 2001). Before 2001, smaller incidents of bacteriological criminal assault, such as intentional contamination of salad bars with Salmonella in Oregon in 1984, and Shigella in muffins and pastries in Texas in 1996 were reported (Chang et al., 2003). A wide range of infectious disease agents have been currently identified that can be intentionally used to cause harm. Diseases caused by these potentially weaponized agents include anthrax, botulism, plague, small pox, tularemia and viral hemorrhagic fevers and have been categorized as ‘Category A’ agents by the CDC (CDC, 2009). Bioterrorism related incidents can have huge economic impacts on the society. Researchers have estimated the economic impact of an anthrax scenario on 100,000 people to be as $ 26.2 billon (Kaufmann et al., 1997). Current human and animal diagnostic methods for infectious microbial agents are generally based on conventional microbiological culturing techniques followed by biochemical identification which are time consuming, requiring 2 to 7 days for confirmation (FDA 2000) and include costly and laborious sample preparation. Hence, developing rapid detection devices have become an absolute necessity to strengthen surveillance systems, to provide early identification in timely manner in cases of exposures thus reducing the possibility of developing symptoms, and also to prevent further transmission of the agents. Biosensors are promising, inexpensive, and efficient alternatives for early identification of such infectious pathogens. Biosensor technology is rapidly evolving in the field of pathogen detection with a wide range of detection platforms being explored recently. The emergence of nanotechnology has further allowed the integration of different nanomaterials into these biodetection systems for developing field-ready, low cost, sensitive, and specific devices for pathogen detection. Different nanostructured materials such as silicon nanowires, carbon nanotubes, gold and magnetic nanoparticles have been employed as transducer materials in biosensor applications. Recently, conductive magnetic nanostructures consisting Of a magnetic core and an electrically active conducting polymer shell have been developed. The conducting polymer, polyaniline, is the most exploited in such magnetic nanostructures. These electrically active polyaniline coated magnetic (EAPM) nanostructures have magnetic properties in addition to the electrical properties, electrochemical activity, mechanical strength, and chemical flexibility of the polymer material. Current literature shows that such novel EAPM nanostructures have not been explored in biosensor detection systems till date. This dissertation describes the development of EAPM nanostructure based detection systems that can utilize the magnetic, electrical, and electrochemical properties of the EAPM nanoparticles for magnetic concentration and biosensor transduction in one system for rapid and specific identification of infectious pathogens. Two different detection approaches have been demonstrated to show the versatility of these novel EAPM nanoparticles in detecting Bacillus anthracis as the model target pathogen. In the first approach, an EAPM based immunosensor was developed that coupled immunomagnetic separation with direct electrical detection of the EAPM nanoparticles in a membrane based charge transfer biosensor. In the second approach, an EAPM based DNA biosensor was developed that involved magnetic concentration of DNA targets combined with direct electrochemical detection of EAPM nanoparticles on a screen printed biosensor. Chapter 2 of this dissertation provides an overview on biosensors and nanotechnology with respect to detection of pathogenic microorganisms and also covers background information on polyaniline and magnetic polyaniline nanostructures. Chapter 3 highlights the novelty and significance of the research conducted and summarizes the research hypothesis and objectives. Chapter 4 walks through the materials and methods involved in synthesizing and characterizing the EAPM nanostructures (Section 4.1), in developing the EAPM based immunosensor and EAPM based DNA biosensor (Sections 4.2 and 4.4), and in implementing the biosensors for detection (Sections 4.3 and 4.5). Chapter 5 presents the results obtained from the objectives set for this research. The first part of this chapter (Section 5.1) discusses the different characterization studies performed on the synthesized EAPM nanostructures. The second part (Sections 5.2 and 5.3) discusses fabrication and performance of the EAPM based immunosensor in pure samples and complex matrices. The final part (Sections 5.4 and 5.5) describes the fabrication and performance of the EAPM based DNA sensor in PCR amplified DNA targets. This is followed by the concluding remarks and future work in Chapter 6. Finally, the appendices give statistical analysis results and data from certain experimental studies. CHAPTER 2: LITERATURE REVIEW 2.1 BA CILL US ANT HRA CIS AND ANTHRAX Bacillus anthracis is a gram-positive, non-motile, aerobic, facultative anaerobic, spore-forming, rod-shaped bacterium and is the etiological agent of anthrax. Anthrax is primarily a zoonotic disease but all mammals, particularly humans, may develop this disease. The spore forms of Bacillus anthracis are highly resistant to environmental conditions such as heat, ultraviolet and ionizing radiation, pressure and chemical agents. They are able to survive for long periods of time in contaminated soils and this account for the ecological cycle of the microorganism. However, it is not clear whether the complete life cycle from germination to sporulation occurs outside the host (Mock and Fouet, 2001). The vegetative cells of the bacterium are square ended and capsulated having a size range of 3 to 5 um while the spores are elliptical with a size range of 1 to 2 pm. 2.1.1 Virulence of Anthrax The primary Virulence factors of B. anthracis are toxin production and capsule formation. Virulent strains of the microorganism carry two large plasmids pXOl and pX02 which encode these virulence factors. The plasmid pXOl carriesthe structural genes for the anthrax toxin proteins pagA (protective antigen), lef (lethal factor), and cya (edema factor); two trans-acting regulatory genes atxA and pagR; a gene encoding type I topoisomerase, topA; and a three gene operon, gerX, which affects germination. Plasmid pX02 carries three genes which encode capsule synthesis, capB, capC, and capA; a gene associated with capsule degradation, dep; and a trans-acting regulatory gene acpA (Okinaka et al., 1999). Figure 2-1 shows a schematic representation of the B. anthracis plasmids and their associated virulence. bicarbonate pXOl 4,, / temperature k/ V~ I dep capA capC capB lef pagR pagA gerX atx “ ® 69 Figure 2-1 Plasmids and genes associated with virulence of B. anthracis (adapted from Mock and Fouet, 2001). B. anthracis toxins are composed of three proteins (Figure 2-2): protective antigen (PA), lethal factor (LF) and edema factor (EF). The mature PA protein, an 83 kDa protein, is made up of 735 amino acids (Petosa et al., 1997). The mature LP (90 kDa) is a protein that contains 776 amino acids (Pannifer et al., 2001). The mature EF protein (89 kDa) is an inactive enzyme that has 767 amino acids (Leppla, 1982). None of the above proteins are toxic separately. Toxicity is associated with the formation of binary exotoxins. The association of PA and LF results in the formation of lethal toxin (LTx), which provokes lethal shock in animals, while the association of PA and EF forms the edema toxin (ETx), which produces edema in the skin (Collier and Young, 2003). Lethal factor Protective antigen Edema factor Lethal Toxin (LTx) Edema Toxin (ETx) Lethal shock Edema Figure 2-2. Mode of action of B. anthracis toxin proteins. The Bacillus anthracis capsule is a polymer of y-D-glutamic acid. The molecular weight of the polyglutamic chains is in between 20 and 55 kDa in vitro and estimated to be 215 kDa in Vivo. The capsule enables the bacteria to evade the host’s immune-system and provoke septicemia thus resulting in pathogenicity. However, the capsule by itself is a monotonous linear polymer which is weakly immunogenic and does not favor immune response (Mock and Fouet, 2001). 2.1.2 MS of Infection Anthrax is initiated by the entry of the spores into the host body through skin lesions, insect bite, ingestion of contaminated meat or inhalation of airborne spores. Depending on the route of infection, there are three clinical forms of anthrax: cutaneous, gastrointestinal and inhalational (pulmonary). The cutaneous form accounts for 90% of all human cases (Spencer, 2003) and occurs when the spores enter the host body through a skin lesion and appears at first as a small pimple developing into a painless back eschar within few days. Antibiotic treatment is highly effective in cases of cutaneous anthrax which has a mortality rate of about 20% in untreated cases (CDC, 2001). The gastrointestinal form of anthrax is extremely rare but potentially fatal. It frequently occurs after the ingestion of undercooked meat from infected animals and is characterized by fever, nausea, vomiting, abdominal pain and bloody diarrhea (Mock and Mignot, 2003). Gastrointestinal anthrax has been reported to cause fatalities in 25-60% of cases (CDC, 2001). The inhalational form of anthrax is considered the most dangerous among the three having a mortality rate nearing 100% (CDC, 2001). The inhaled spores reach the alveolus where they are phagocytosed by macrophages and transported to the mediastinal lymph nodes, where spore germination can occur up to 60 days. The disease progresses rapidly following germination resulting in the production of exotoxins that cause edema, necrosis, and hemorrhage (Mock and Fouet, 2001). Diagnosis is difficult in both gastrointestinal and inhalational forms resulting in the disease to become treatment resistant and rapidly fatal. 2.1.3 Anthrax and Biosecurity The use of microorganisms as biological weapons has long been reported in history. One of the first major attacks that have been reported occurred in the 14th century with Yersenia pestis during the siege of Kaffa (Inglesby et al., 2000). The most recent was the deliberate release of B. anthracis spores through the postal system in the United States in October 2001, shortly after the September 2001 terrorist attack in New York and Washington resulting in 22 cases of anthrax and 5 deaths (Jemigan et al., 2001). An optimal biological weapon should be highly lethal, easily produced in large quantities, environmentally stable and have the capability to be readily dispersed into aerosol (i.e. 1 to lOum particle size) (Peruski and Peruski, 2003). When reviewed for these characteristics, B. anthracis is the most likely to be used as a biological weapon since it has the ability to form spores which can be easily aerosolized, inhalational anthrax has a high mortality rate nearing 100%, and the spore forms of the bacteria are very stable surviving harsh environmental conditions. It has been estimated that the release of 50 kg of dried anthrax spores for two hours can lead to a complete breakdown in medical resources and civilian infrastructure in a city of 500,000 inhabitants (Spencer, 2003). The Centers for Disease Control and Prevention (CDC) has categorized B. anthracis as a category A agent or a high priority agent. Category A agents include agents that can be easily transmitted from person to person, cause high mortality with potential for major public health impact, may cause public panic and social disruption, and require special action for public health preparedness (CDC website). 2.1.4 Traditional Identification Methods Detection and specific identification of B. anthracis require complex techniques and laborious methods because of the genetic similarities among various Bacillus species as well as their existence in both spore forms and vegetative state. B. anthracis is identified using standard biochemical techniques such as its sensitivity to penicillin, non-motility, non B-hemolytic behavior on sheep or horse blood agar plates and its susceptibility to lysis by gamma phage. It has been reported that identification of B. anthracis by initial blood culturing requires 6-24h for growth, which is followed by morphological and biochemical identification that requires additional 12-24h and finally definitive identification that requires an additional 1-2 days (Inglesby, 2000). B. anthracis is also shown to selectively grow on polymyxin-lysozyme EDTA-thallous acetate (PLET) agar which requires 1-2 days for grth followed. by further confirmation (Erickson and Kornacki, 2003). 2.2 BIOSENSORS IN PATHOGEN DETECTION 2.2.1 Introduction A biosensor can be defined as an analytical device that integrates a biological sensing element with an electrical transducer to quantify a biological event (for e.g. an antigen-antibody reaction) into an electrical output. The basic concept of operation of a biosensor has been illustrated in Figure 2-3. The biological sensing elements include enzymes, antibodies, DNA, aptamers, molecularly imprinted polymers, microorganisms and whole cells. Depending on the transducing mechanism used, biosensors can be of many types such as: electrochemical, electrical, optical, mechanical and magnetic biosensors. The different transducing mechanisms and biological sensing elements are discussed in detail in later part of this section. ANALYTE BIORECEPTOR TRANSDUCER SIGNAL PROCESSING ' “G Antibodies Electrochemical Nucleic Acids Electrical 0 «<1 Ottaal :> l:> MESS... Enzymes Mechanical . «C Whole Cells Magnetic Figure 2-3. Schematic representation of a biosensor. Biosensor technology is emerging as a promising field for rapid detection of pathogens. Currently, the two most popular methods for detection of pathogens are: microbial culturing followed by biochemical identification, and polymerase chain reaction (PCR) assay. The main disadvantages of the conventional microbial culturing techniques are the time consuming process and the multistep assay procedures for food samples requiring pre-enrichment steps. The PCR technology, although highly sensitive, 10 has some disadvantages, such as requirement of expensive equipments, skilled personal to perform assays, DNA extraction stages which increase the detection time, and prior information of target DNA sequences. On the other hand, biosensors show high sensitivity and specificity to targets and can be used as simple one-shot measurement tools or as multi-measurement devices. Moreover, biosensors can be designed to be operated on a real-time basis thus eliminating the need of expensive lab—based testing (Terry et al., 2005). The miniaturization ability of biosensors and their compatibility with data processing technologies allow them to be integrated into small portable devices. This versatility in biosensors has prompted worldwide research and commercial exploitation of the technology. Recent trends indicate (Figure 2-4) that biosensors is the fastest growing technology for rapid detection of pathogens (Lazcka et al., 2007). 120 . 1 5 Forecast 1 I a 100* 5 / PCR m . E 1 E 3 . -2 801 8 i . / Culture Methods '5 604 . Biosensors E . \ 2 / \l 8 40‘ a l - 20‘ / - ————» ELISA f . ' VA/ _'_“"" Gel Electrophoresis 0+}??? _I ff, I ' I ' i I f . I 1985 1990 1995 2000 2005 2010 Figure 2-4. Recent trends in pathogen detection (adapted from Lazcka et al., 2007). 11 2.2.2 Biosergor Transducing Mechanisms 2. 2. 2.1 Mechanical Biosensors 2.2.2.1.] Quartz Crystal Microbalance (OCM) Biosensors Quartz crystal resonators form the basis of QCM sensors. The term “QCM” is used collectively for bulk acoustic wave (BAW), quartz crystal resonance sensors (QCRS) and thickness shear mode (TSM) acoustic sensors (Cooper and Singleton, 2007). QCM sensors comprise Of a thin quartz disc with electrodes plated on it. When an oscillating electric field is applied across the disc, an acoustic wave with a certain resonant frequency is induced. The disc can be coated with a sensing layer of biomolecules based on the analyte to be detected. The interaction of the analyte with the biomolecules on the disc surface causes a change in mass and a concurrent change in resonance frequency that can be directly correlated to the biomolecular interactions(O'Sullivan and Guilbault, 1999). The relation between mass and the resonant frequency is given by the Sauerbrey equation: —2.3x106F2Am AF: 0 A (2.1) where, AF is the change in frequency (Hertz), F 0 is the resonant frequency of the crystal (MHz), A m is the deposited mass (grams) and A is the coated area (cmz). The quartz crystals are inexpensive, easily available and robust thus making them suitable for chemical sensors and biosensors. In addition, QCM based sensors provide great flexibility, wide dynamic range of frequency measurements and label free detection (O'Sullivan and Guilbault, 1999). 12 A wide range of non-labeled QCM biosensors have been reported in the literature for the detection of pathogenic bacteria and viruses. QCM sensors based on lectin recognition systems for bacterial identification have been studied by Shen et al. and Safina et al. (Shen et al., 2007;Safina et al., 2008). Shen et al have used a combination of mannose self-assembled monolayer (SAM) and lectin concanavalin A for the detection of E. coli W1485 in a linear range of 7.5 X 102 to 7.5 x 107 cells/m1. Safina et al. utilized lectin reporters to develop a flow injection QCM biosensor for detection of Campylobacterjejuni and Helicobacter pylori. The authors were able to detect 103 to 105 cells/ml in 30 min. A SAM based QCM immunosensor was developed for the detection of E. coli 0157: H7 by Su et al. (Su and Li, 2004a). The immunosensor was able to detect the target bacteria in the range of 103 to 105 CFU/ml in 30-50 min. Detection of B. subtilis spores as a surrogate to B. anthracis was achieved by Lee et al. utilizing a QCM immunosensor to a detection limit of 450 spores/ml (Lee et al., 2005). Furthermore, virus (dengue virus and hepatitis B virus) detection with QCM immune and nucleic acid based sensors have been reported by Wu et al. and Yao et al. (Yao et al. , 2008;Wu et al., 2005) QCM biosensors for the detection of DNA sequences have also been developed using nanoparticle labels as amplifiers. Mao et al. reported the use of streptavidin conjugated Fe3O4 nanoparticles for the detection of E. coli 0157: H7 eaeA gene. The nanoparticles acted as ‘mass enhancers’ and amplified the change in frequency. The biosensor could attain a sensitivity of 1012 M synthetic oligonucleotides and 2.67 X 102 CFU/ml E. coli 0157: H7 cells (Mao et al., 2006). Similarly, Au nanoparticles were employed by Wang et al. for real time bacterial DNA detection in a circulating flow 13 QCM biosensor. The authors reported a sensitivity of 2.0 X 103 CFU/ml for E. coli 0157: H7 eaeA gene (Wang et al., 2008). Although QCM biosensors are being used widely for pathogen detection, they have limited potential for scalability since the mass detection capability scales with the QCM sensor surface. 2. 2. 2. 1.2 Surface Acoustic Wave (SA W) Biosensors SAW sensors are the second class of acoustic wave sensors that have found applications in biosensor devices. SAW sensors consist of two metal interdigital transducers (IDT) etched from a thin metal film deposited on a piezo-electric substrate. The sensing mechanism is based on the changes in SAW velocity or attenuation when mass is sorbed on the sensor surface. Since the acoustic energy is strongly confined to the surface, SAW devices are very sensitive to changes in the surface such as mass loading, viscosity, and conductivity changes (Galipeau et al., 1997). It has been suggested that SAW based biosensors have greater sensitivities than QCMS because of their higher mass sensitivities (~ 200 times > QCM) (Galipeau et al., 1997). SAW biosensors have been successfully applied for the detection of bacteria and viruses. E. coli detection using SAW biosensors have been reported by multiple authors in the literature (Berkenpas et al., 2006;Deobagkar et al., 2005;Moll et al., 2007;Moll et al., 2008). The biosensors have used antibodies as the biological sensing element with sensitivities ranging from 106 cells/ml to 0.4 cells/pl. Branch and Brozik have developed a 36° YX cut LiTaO3 based SAW device for the detection of the Bacillus anthracis simulant Bacillus thuringiensis spores in aqueous conditions (Branch and Brozik, 2004). The authors have investigated two waveguide materials polyimide and polystyrene for 14 creating the Love wave sensors. Detection of Bacillus thuringiensis spores at concentrations below the lethal dose of anthrax spores was possible using both waveguide materials. The sensor had a detection limit of a few hundred cells per ml and a response time of < 100 s. Jin et al. developed a SAW biosensor for detecting the gene of Staphylococcal Enterotoxin B utilizing ST-cut quartz and SiOz guiding layer. The biosensor had a sensitivity of 10 ng/ml and a linear range of 35-200 ng/ml (Jin et al., 2003). Recently, SAW biosensors were used for detecting viral bioagents by Bisoffi et al. (Bisoffi et al., 2008). A lithium-tantalate based SAW transducer with SiOz waveguide sensor platform was used for Coxsackie virus B4 and Sin Nomber Virus detection. SAW resonators are suitable for use in simple electronic setups because of their low insertion losses and sharp resonance frequencies. As a result, insertion of such devices into oscillator circuits is beneficial as such circuits are commonly used in point-of—care diagnostics. Furthermore, the SAW based biosensors can be prepared from cheap components thus making them suitable for integration into inexpensive sensor arrays (Lange et al., 2008). 2. 2. 2. 1.3 Microcantilever Based Biosensors Microcantilever based biosensors are derived from microfabricated cantilevers used in atomic force microscopy (AFM) and are based on the bending induced in the cantilever when a biomolecular interaction takes place on one of its surface which is translated into nanomechanical motion and is commonly coupled to an optical or piezoelectric readout system (Carrascosa et al., 2006). The cantilevers can be operated in the static deflection mode where analyte binding causes cantilever bending or in the dynamic resonant mode where analyte binding causes change in resonant frequency 15 (Waggoner and Craighead, 2007). Microcantilever sensors are promising for biosensor applications since they can perform local, high resolution, and label-free molecular recognition measurements (Carrascosa et al. , 2006). Davila et al. have demonstrated microcantilever based biosensors in the detection of B. anthracis Sterne strain in air and water (Davila et al., 2007). The detection scheme involved measurement of the decrease in resonant frequency driven by thermally induced oscillations as a result of the mass of spores measured by a laser Doppler vibrometer. The authors reported a minimum detection of 2 spores (740 fg) and 50 spores (139 pg) in air and water, respectively, using 20 um long, 9 pm wide, and 200 nm thick cantilevers. Muthasaran and coworkers have utilized piezoelectric-excited millimeter-sized cantilever sensors for the detection of B. anthracis Sterne spores and E. coli 0157: H7 cells. The sensors consisted of a piezoelectric and a glass layer and were able to detect B. anthracis spores at 300 spores/ml and E. coli 0157: H7 cells in ground beef at 50 to 100 cells/ml. Extremely sensitive microcantilever based biosensors capable of detecting a single pathogen have also been reported in literature(Campbell and Mutharasan, 2006). Illic et al. were able to detect a single E. coli 0157: H7 cell using low stress silicon-nitride cantilever beams in air (Ilic et al., 2001). The mass of a single E. coli 0157: H7 cell was found by the authors to be 665 fg. Similarly, Johnson et al. reported the use of microscale silicon cantilever resonators for vaccinia virus detection in air (Johnson et al. , 2006). The authors measured the mass of a single vaccinia virus particle to be 12.4 21:13 fg and 7.9 i: 4.6 fg using two different sized cantilever beams. 16 2. 2.2.2 Optical Biosensors 2. 2. 2. 2.1 Surface Plasmon Resonance (SPR) Biosensors SPR is an Optical technique for monitoring biomolecular interactions that occur in the close vicinity of a transducer surface. SPR based biosensing can be subdivided into three categories depending on the mode of SPR detection: angular SPR biosensing, spectral SPR biosensing, and local SPR biosensing. Angular SPR biosensing is the most common form of SPR biosensing and involves attenuated total reflection approach using Kretschmann geometry (Erickson et al., 2008). Spectral SPR biosensing is conducted at a fixed incident angle and utilizes the wavelength dependence of the dielectric constant of the metal film to interrogate the surface plasmon coupling conditions. Local SPR (LSPR) biosensing or nanoparticle based SPR involve coupling of surface immobilized metallic nanoparticles or nanostructures into a plasmon mode which results in a decrease in the transmitted power at a specific resonant wavelength dependant on environmental dielectric conditions (Erickson et al. , 2008). SPR biosensors provide several advantages over conventional transduction techniques, such as, capability of label-free detection, ability to produce continuous real-time responses, regeneration of the active sensor surface, feasibility for miniaturization, sensitive detection of small molecules, and multiplexing ability ((Shankaran et al., 2007). Particularly, multiplex detection, or detection of different targets on the same platform, is an important characteristic in these biosensors since it adapts to screening of unknown samples. SPR based immunosensors have been developed for the detection of E. coli 0157: H7 by Irudayaraj and coworkers. The authors have demonstrated a sensitivity Of 103 17 CFU/ml for the pathogen using a SAM based SPR biosensor and the commercially available Spreeta SPR biosensor (Waswa et al., 2007;Subrarnanian et al., 2006). Detection of Salmonella Typhimurium in chicken carcasses was achieved using an antibody based SPR biosensor by Lan et al. (Lan et al., 2008). The SPR biosensor had a lowest detection limit Of l X 106 CFU/ml. Highly sensitive detection of Salmonella Enteritidis was attained by Waswa et al. using the commercial BiacoreTM SPR biosensor (Waswa et al., 2006). The limit of detection (LOD) of the biosensor as reported by the authors was 23 CFU/ml for Salmonella. Chen et al. have demonstrated an immunomagnetic separation based SPR detection method for the foodbome pathogen Staphylococcus aureus (Chen et al., 2007). The detection system involved an initial immunomagnetic bead based separation step of 30 min and had a sensitivity of 106 CF U/ml in a total assay time of 2 h. A SAM based SPR immunosensor was developed by Jyoung et al. for the detection of Vibrio cholerae 01 with a detection range of 105 to 109 cells/m1 (.1 young et al. , 2006). Detection of viruses using SPR based biosensors have been reported by Chung et al. and Vaisocherova et al. (Chung et al. , 2005;Vaisocherova et al. , 2007). Vaisocherova and coworkers developed a SPR biosensor for detecting antibodies against Epstein-Barr Virus. The antibody detection was performed using an immunoreaction between the antibody and a synthetic peptide for the Virus. The sensor had a sensitivity of 0.2 ng/ml (~ lpM). Furthermore, multiplex detection of four foodbome bacterial pathogens (E. coli 0157: H7, Salmonella Typhimurium, Listeria monocytogenes & Campylobacter jejuni) was demonstrated by Taylor et al. using an eight-channel SPR biosensor (Taylor et al., 2006). The LOD for each Of the four species Of bacteria was in the range of 3.4 X 103 and 1.2 X 105 CFU/ml. 18 Several detailed review articles are available in literature that focus on different SPR based detection techniques and their applications (Shankaran et al., 2007;Homola, 2008;Hoa et al. , 2007). 2. 2. 2. 2.2 Fluorescence Based Biosensors Fluorescence is the radiative deexcitation of a molecule following the absorption of a photon. Generally, the emitted photon is of lower energy than the absorbed photon, and the fluorescence emission peak of a species is at longer wavelength than the absorption peak, the wavelength separation being referred as Stoke’s shift. Fluorescence based detection systems have gained popularity in biosensors due to their high sensitivity and are mostly based on the detection of the fluorescent signal generated by fluorophores used to label the biomolecules. Fluorescence detection techniques can be performed in the high-throughput mode in combination with platforms such as microarrays. Microarrays offer the advantage of using a two-dimensional layout of recognition elements for simultaneous detection and quantification. Taitt et al. have demonstrated a fluorescence based microarray immunosensor for the simultaneous detection of nine targets comprising of B. anthracis Sterne, B. globigii, Francisella tularensis, Yersenia pestis, S. Typhimurim, Staphylococcal enterotoxin B, ricin, cholera toxin and M82 coliphage (Taitt et al., 2002). Li et al. have developed DNA-based fluorescence nanobarcodes to integrate with DNA- microarrays for the simultaneous detection of four targets (B. anthracis, Francisella tularensis, Ebola virus and SARS coronavirus) (Li et al., 2005). The feasibility study involved confocal microscopy, dot blotting and flow cytometry which resulted in attomolar sensitivity. 19 Fluorescence resonance energy transfer (FRET) based detection involves non- radiative energy transfer between a donor fluorophore and an acceptor fluorophore when they are in close proximity (Epstein et al., 2002). A fiber optic portable biosensor utilizing the principle of FRET was developed by K0 and Grant for rapid detection of S. Typhimurium in ground pork samples (K0 and Grant, 2006). The biosensor had a sensitivity of 105 CFU/ml in a response time of 5 min. Kim et al. reported a molecular beacon DNA microarray system for fast detection of E. coli 0157: H7 based on FRET (Kim et al., 2007a). In this system, unlike conventional fluorophore-quencher beacon design, two fluorescence molecules allowed active visualization of both hybridized and unhybridized states of the beacon. The target gene detection limit for the system was lng/ul. Fluorescence based tapered fiber optic biosensors have also been employed in pathogen detection. A fluorescence based fiber optic biosensor for detecting E. coli 0157: H7 in ground beef samples was developed by Geng et al. (Geng et al., 2006). The authors reported sensitivity of 103 CFU/ml in pure cultures and of 1 CFU/ml in artificially contaminated ground beef samples after 4 h enrichment using a sandwich immunoassay. Nanduri et al. developed an automated fiber optic based immunosensor called RAPTORTM for the detection of Lysteria monocytogenes in food samples. The LOD of the system was 5 X 105 CFU/ml in food samples and 1 x 103 CFU/ml in PBS (Nanduri et al., 2006). Inspite of the enhanced sensitivity, all fluorescence based biosensors suffer from disadvantages attributed mainly to the characteristics of the fluorescent probe such as 20 stability and efficiency of labeling. The expensive fluorescent labels are also a drawback of this technology. 2. 2.2.3 Electrochemical Biosensors Electrochemical biosensors are based on the detection of electrochemical signal generated by consumption or production of electrons from biological interactions occurring at the sensor surface. Advantages such as low cost, high sensitivity, miniaturization ability, low power requirements, and simple instrumentation make the electrochemical biosensors well suited for clinical and environmental analysis. Electrochemical biosensors are generally classified as amperometric, potentiometric, conductometric and impedimetric. 2. 2. 2. 3. I Amperometric Biosenm Amperometric biosensors are based on the measurement of current changes resulting from oxidation or reduction of an electroactive species in a biochemical reaction. The current is typically measured at a fixed potential (amperometry) or during controlled variations of the potential (voltammetry). Theegala et al. reported an oxygen-electrode based amperometric biosensor for the qualitative detection of E. coli 0157: H7 in water (Theegala et al., 2008). The biosensor detected changes in oxygen concentration due to decrease in enzymatic activity upon binding of bacterial cells. The biosensor could detect as low as 50 cells/ml in 20 minutes. A renewable amperometric immunosensor for the detection of S. Typhi was reported by Singh et al. (Singh et al., 2005). The detection technique involved a sandwich ELISA system with a LOD of 105 cells/ml in 90 minutes. Amperometric detection of antibodies against Bacillus anthracis protective antigen was also achieved by Aguilar et al. (Aguilar 21 and Sirisena, 2007). The antibodies were captured and detected using microcavities with a LOD of 10 fg in a 200 nL sample. A disposable amperometric immunosensor based on screen printed electrode (SPE) coated with agarose/nano-Au membrane and horseradish peroxidase labeled antibody for specific detection of the foodbome pathogen Vibrio parahaemolyticus was developed by Zhao et al. (Zhao et al., 2007). The immunosensor showed a sensitivity of 7 X 104 CFU/ml for the pathogen and its accuracy was in good agreement (97.5%) with ELISA results. Lermo et al. described a genomagnetic assay for the electrochemical detection of Salmonella spp. based on in situ DNA amplification and magnetic primers (Lermo et al., 2007). Detection was achieved by sandwich hybridization of the target on magnetic beads which were then separated by a magneto electrode based on graphite—epoxy composite followed by electrochemical detection using an enzyme marker anti-digoxigenin horseradish peroxidase. The authors achieved a sensitivity of 2.8 frnol with PCR amplicons. Multianalyte detection using electrochemical genosensors have also been studied by Farabullini et al. and Elsholz et al. (Elsholz et al., 2006;Farabullini et al., 2007). F arabullini et al. achieved nanomolar detection limits for the pathogenic bacteria, Salmonella sp., E. coli 01572H7, L. monocytogenes and S. aureus using differential pulse voltammetry to detect u-napthol signal in less than 1 h. 2. 2. 2. 3. 2 Potentiometric Biosen_s_0_rs Potentiometric devices are based on the measurement of accumulation of charge potential at the working electrode of an electrochemical cell in comparison to the reference electrode with zero or negligible current flow between the electrodes. The 22 measured potential is related to the concentration of the analyte through the Nersnt equation: E = 150 i (RT/nF)1nQ (2.2) where, E is the cell potential at zero current, E0 is the standard potential, R is the universal gas constant, T is the absolute temperature, F is the Faraday constant, n is the total number of charges of ion, Q is the ratio of ion concentration at the anode to that at the cathode (Diamond, 1998;Eggins, 2002). Among electrochemical transducing methods, potentiometric methods are the least exploited in pathogen detection due to their high detection limits and poor selectivity, the main advantage of these devices being wide detectable concentration range and continuous measurement capability (Palchetti and Mascini, 2008). Another approach involves ion selective field effect transistors (ISFETs) that employ semiconductor field- effect to detect biorecognition events. However, the application of these devices in biosensors has been limited by production problems related to immobilization, fabrication and packaging, poor detection limits and device stability (Lazcka et al. , 2007). An advancement that has evolved from the ISFET is the light addressable potentiometric sensor (LAPS) which combines potentiometry with optical detection (Hafeman et al., 1988;Lazcka et al., 2007). Ercole et al. reported an antibody based LAPS biosensor for determination of E. coli in food (Ercole et al., 2003). The biosensor detected variations in pH due to ammonia production by urease-E. coli antibody conjugates in commercial lettuce, Sliced carrot, and rucola samples. The sensor was able to reach a sensitivity of 10 cells per ml in an assay time of 1.5 h. 23 2. 2. 2. 3. 3 Conductometric Biosensors Conductometric biosensors utilize the electrical conductivity of a sample to determine the components and their concentration (Rahman et al., 2008). Muhammad Tahir and Alocilja (Muhammad-Tahir and Alocilja, 2003a;Muhammad-Tahir et al., 2005) have developed a conductometric biosensor for the detection of pathogenic bacteria and viruses. The biosensor was fabricated using conducting polyaniline as an electronic label in a sandwich immunoassay scheme and the authors demonstrated that polyaniline improved the sensitivity of the biosensor by forming a conductive molecular bridge between silver electrodes. The authors reported a sensitivity of 8.3 X 101 CFU/ml for Salmonella, 7.9 x 101 CFU/ml for E. coli 01572H7, 7.5 x 101 CFU/ml for E. coli and 103 CCID/ml for BVDV virus in a detection time of 10 min with the conductometric biosensor. A conductometric immunosensor based on magnetic nanoparticles has been recently developed for the detection of E. coli by Hnaiein et al. (Hnaiein et al., 2008). The immunosensor was composed of streptavidin modified magnetic nanoparticle layer immobilized on a conductometric transducer consisting of interdigitated gold electrodes. Conductivity measurements allowed detection of 0.5 CFU/ml of E. coli without the need for amplification. 2. 2. 2. 3. 4 Impedimetric Biosensors Impedance spectroscopy involves applying small amplitude perturbing sinusoidal voltage signal to an electrochemical cell and measuring the resulting current response. The complex impedance, sum of real and imaginary impedance components, can be calculated as a function of the excitation frequency of the applied potential by varying it over a range of frequencies (Katz and Willner, 2003). Impedimetric detection techniques 24 provide advantages of high sensitivity, linearized current-potential characteristics, measurement over wide time or frequency range and label free sensing (Rahman et al., 2008;Lazcka et al. , 2007) Radke and Alocilja had developed a microimpedance biosensor for the detection of E. coli (Radke and Alocilja, 2005). The sensor detected changes in impedance caused by the presence of bacteria immobilized on interdigitated gold electrode arrays fabricated from silicon. The biosensor was able to discriminate between different cellular concentrations of the bacteria (105 to 107 CFU/ml) in 5 min. Nandakumar et al. have demonstrated the detection of S. Typhimurium using electrochemical impedance spectroscopy based on Bayesian decision theory (Nandakumar et al., 2008). The technique detected the pathogen in 6 min at a lowest concentration of 500 CFU/ml. An impedance biosensor based on interdigitated array microelectrode coupled with magnetic nanoparticle- antibody conjugates was developed for rapid and specific detection of E. coli 01 57:H7 in ground beef samples by Varshney et al. (Varshney and Li, 2007). Magnitude of impedance and phase angle was measured in a frequency range of 10 Hz to 1 MHz in the presence of 0.1 M mannitol solution. The lowest detection limit of the biosensor for E. coli OlS7:H7 was 7.4 X 104 CFU/ml in pure cultures and 8.0 X 105 CFU/ml in ground beef samples, the total detection time being 35 min. 2. 2. 2.4 Magnetic Biosensors Devices based on the detection of magnetic labels are emerging as a promising new approach in the field of biosensing. Magnetic labels have gained popularity in biosensing because they are physically and chemically stable, are relatively inexpensive, and can be easily made biocompatible. Several approaches have been developed in the past few 25 years for both direct and indirect detection of magnetic labels. Direct detection includes approaches for measuring magnetic parameters such as magnetic permeability, magnetic remanence, magnetoresistance, and Hall Effect. The indirect detection methods are based on micro-cantilever based force amplified sensors and magnetic relaxation switches (Tamanaha et al., 2008). However, the applications of these magnetic devices for detection of actual targets such as pathogenic microorganisms remain limited and require further research. Edelstein et al. had developed a multi-analyte BARC (Bead Array Counter) biosensor using giant magnetoresistive (GMR) sensors to detect and identify biological warfare agents (Edelstein et al., 2000). The prototype designed by the authors consisted of a microfabricated chip with GMR sensor arrays, an electronic chip-carrier board, a fluidics cell and an electromagnet. DNA probes were patterned onto the GMR sensor chips and hybridized with complementary PCR products. Micron sized magnetic beads were then bound to the DNA sample by streptavidin biotin interactions and the unbound beads were removed by applying a magnetic field. The bound magnetic beads were then detected by the GMR sensors. The authors were able to demonstrate the detection of B. anthracis lethal factor and C. botulinum neurotoxin A using the BARC biosensor. - A mass-sensitive magnetoelastic immunosensor for the detection of E. coli 0157:H7 was reported by Ruan et al. (Ruan et al., 2003). The detection was based on the immobilization of alkaline phosphatase labeled antibodies on the surface of a micrometer-scale magnetoelastic cantilever and amplification of the mass change associated with antigen-antibody binding reaction by biocatalytic precipitation of bromo- 26 4-chloro-3 -indolyl phosphate. The minimum detectable level of the immunosensor was 6 X 102 cells/ml. A high efficiency Hall effect micro-biosensor platform has recently been developed for the detection of magnetically labeled biomolecules by Sandhu et al. (Sandhu et al., 2007). In this system, the integration of Hall-effect structures with micro-current lines allowed manipulation of the magnetic beads position via field gradients. The authors studied the hybridization of fully-complementary DNA strands of 20-25 bases using Dynabeads as magnetic labels with this platform. Although, the sensitivity of the sensor was not reported, the authors were able to demonstrate a quantitative relationship between the number of magnetic labels and the output signal. 2.2.3 Biological Sensing Elements The main classes of biological sensing elements that are currently used in biosensors for pathogen detection are: (i) enzymes, (ii) antibodies, (iii) nucleic acids, (iv) aptamers and (v) molecularly imprinted polymers. Enzymes are large complex macromolecules consisting mainly of protein, and containing a prosthetic group, that often include one or more metal atoms. Enzymes act as biocatalysts in biochemical reactions. Enzymes are mostly used for labeling other biomolecules such as antibodies and DNA in pathogen biosensing similar to assays such as ELISA (Lazcka et al., 2007). Antibodies are molecules consisting of polypeptide chains and can be classified as polyclonal, monoclonal and recombinant antibodies. Antibodies have been extensively applied in biosensors for the following advantages: they are very selective for a particular antigen, they are ultra-sensitive, and can bind very strongly with the corresponding 27 antigen. Antibodies can also be easily labeled with different molecules such as enzymes, radioisotopes, fluorescent probes, chemiluminescent probes or metal tags (Eggins, 2002). Most antibody based biosensors (immunosensors) are based on a competitive or sandwich assay when applied to detection of low and high molecular weight molecules, respectively (Marquette and Blum, 2006). Nucleic acid (DNA & RNA) hybridization is a thermodynamically favored process which is triggered by highly specific interactions between the base-pairs where each nucleotide base strongly binds to its complementary base through multiple hydrogen bonds. This property makes nucleic acids an excellent choice for use as bio-recognition elements in biosensors. The nucleic acid based biosensors reported in literature can be classified based on optical, mass sensitive or electrochemical detection platforms (Diamond, 1998). Of these, DNA-based electrochemical sensors which rely on the conversion of DNA base-pair recognition event into useful electrical signal have been exploited the most. Such DNA based electrochemical devices have the advantages of high sensitivity, selectivity, low-cost, miniaturization ability and minimal power requirements (Wang, 2002). Numerous approaches for electrochemical detection of DNA have been developed which include direct electrochemistry of DNA, electrochemistry at polymer- modified electrodes, electrochemistry of DNA-specific redox reporters, electrochemical amplification with nanoparticles and devices based on DNA-mediated charge transport chemistry (Drummond et al., 2003). Aptamers are synthetic folded DNA or RNA oligonucleotide sequences specifically generated complementary to diverse target molecules. Aptamers offer multiple advantages as bio-recognition elements in biosensors in comparison to natural receptors 28 such as antibodies and enzymes. These include: high selectivity for targets (similar to monoclonal antibodies), smaller size, high reproducibility with minimum batch to batch variations, cost-effectiveness, chemical stability, greater flexibility for biosensor design and synthesizing ability for any given target (Song et al., 2008;Mairal et al., 2008). Molecular imprinting is a template-induced process for the formation of specific recognition sites in a polymeric material, where the template directs the positioning and orientation of the material’s structural components by a self-assembling mechanism. Molecularly imprinted polymers (MIPS) have several advantages over natural biomolecules such as antibodies and receptors. These include their unique stability in harsh environmental conditions, high affinity and selectivity similar to that of natural receptors, and their simplicity of preparation and ease of adaptation to practical applications (Ye and Haupt, 2004;Piletsky et al., 2006). MIPS have been generated for a broad range of molecules from small molecules such as drugs to large proteins and cells, the best results being obtained for molecular weights in the range of 200-1200 Da (Piletsky et al., 2006). 29 2.3 NANOBIOSENSORS & NAN OSTRUCTURED TRANSDUCERS Nanotechnology is defined as the creation of functional materials, devices and systems through the control of matter at the l-100 nm scale (Wang, 2005). Recent advances in nanotechnology have led to the development of a variety of nanomaterials, e.g., nanowires, nanotubes, nanoparticles, nanospheres and nanorods. Due to their extremely small size, these nanomaterials exhibit unique properties such as physical strength, chemical reactivity, electrical conductivity, magnetism and optical characteristics. Such characteristics of the nanomaterials offer enormous prospects in designing novel methods of signal transduction and in integrating them into biosensing devices. Nanomaterials and nanoelectrodes are an important component of nanobiosensors. This section describes different nanostructured transducer materials and their current applications in biosensors with special emphasis on the detection of pathogenic microorganisms. 2.3.1 Nanowire Based Biosensors Recently, nanowires have been explored as biosensor transducers for the following reasons: Firstly, the high aspect ratios of the nanowires make them extremely sensitive to biological interactions. Secondly, the electronically switchable properties of semiconducting nanowires allow direct and label-free electrical detection. Thirdly, the size-tunability of the nanowires to sub-100 nm allows high density device fabrication (Wang et al., 2007a). Semiconducting nanowires including silicon nanowires, and conducting polymer nanowires, as well as metallic nanowires are the most studied ones in biosensor architectures. Laser-assisted metal-catalyzed growth (LCG), Vapor-liquid-solid growth (VLS), Vapor-solid-solid growth (VSS), chemical-solution based growth and 30 template-assisted growth are some techniques employed in the synthesis of nanowires (Spanier, 2006). 2.3.1.] Silicon Nanowires Semiconductor nanowire based detectors that show a change in conductivity with respect to variations in electric potential are referred to as field-effect transistors (F ET). FETs have been widely used as transducers in detection devices because of the high surface to volume ratio and high sensitivity of the carrier mobility to variations in electric field at the semiconductor nanowire surfaces. Silicon (Si) nanowire based FETs are particularly attractive for electrical based sensing because of their superior electrical performance characteristics, high reproducibility and controllability during growth, and their high-performance switching characteristics that affect sensitivity (Cui et al., 2003;Eggins, 2002;Patolsky et al., 2006). In a recent study, Mishra and coworkers have developed a nanowire field effect transistor (nano-FET) for detecting the enterotoxins released by the food borne pathogen Staphylococcus aureus (S. aureus Enterotoxin B, SEB) (Mishra et al., 2008). The nano- F ETs were fabricated using 50 nm doped polysilicon nanowires that were attached to small gold terminals separated by a gap of 150 nm using c beam lithography with controlled reactive ion etching using chlorine plasma. The authors reported a very low sensitivity of 10 1M of pure SEB using electrochemical impedance spectroscopy. Detection of single influenza A viruses was achieved by Patolsky et al. with antibody modified Si nanowire arrays (Patolsky et al., 2006). The detection was based on the change of conductance of the nanowire device from a baseline value upon binding of the 31 virus particle followed by simultaneous return of the nanowire conductance to the baseline value upon unbinding of the viruses Si nanowires have also been applied for the detection of DNA sequences. A two- terminal Si nanowire based electronic device for ultrasensitive DNA detection have been described by Hahm et al. (Hahm and Lieber, 2004). The device was modified by the authors with peptide nucleic acid (PNA) receptors that could distinguish between the wild type and and AF508 mutation site in the cystic fibrosis trans membrane receptor gene. The time dependant conductance change due to PNA-DNA hybridization demonstrated that the device could selectively detect DNA concentrations as low as 10 M. The detection limit of the device as suggested by the authors were better than other real-time measurement systems such as nanoparticle enhanced surface plasmon resonance (SPR) sensors and quartz crystal microbalance. More recently, arrays of highly ordered n-type silicon nanowires were fabricated using complementary metal-oxide semiconductor (CMOS) technology by Gao et al. (Gao et al., 2007). The Si nanowire arrays enabled real time linear detection of DNA hybridization to its complementary target over a large dynamic range of 25 fM to 5 pM with a detection limit of 10 tm. 2.3. 1.2 Conducting Polymer Nanowires Conducting polymer (CP) nanowires or ‘organic’ nanowires are another class of nanowires that are used as transducer materials in biosensors. Simple synthetic procedures, controllable electrical properties, flexible chemical structures, compatability with biomolecules and efficient charge transfer in biochemical reactions are some of the properties which make CP nanowires attractive candidates for biosensing. Polyaniline and polypyrrole are the two most studied CPS in biosensor applications. 32 Pal et al. have described a polyaniline nanowire based direct-charge transfer biosensor for the detection of Bacillus species (Figure 2-5) (Pal et al., 2008). The biosensor used two sets of antibodies (secondary and capture antibodies) for detection of the food borne pathogen, Bacillus cereus. The detection principle involved an immunoreaction coupled with direct electron charge flow by polyaniline nanowires across silver electrodes. The biosensor could attain sensitivity as low as 10l CFU/ml in 6 min through a reagentless process. The biosensor showed little or no cross-reactivity with other non-target bacteria and was also able to detect the presence of the target in a mixed culture of five different microorganisms. Application Conjugate pad Pad Capture Pad Electrode ’ Absorption .. . i a " Pad .m- at, .. .. . as .l- » .1... . , Platform .3... at: a» Figure 2-5. Schematic of the polyaniline nanowire biosensor for Bacillus cereus detection (Pal et al., 2008). Fan et al. have reported a nanogapped microelectrode-based biosensor array for detection of microRNAs (miRNAs) utilizing polyaniline nanowires (Fan et al., 2007). PNA capture probes were immobilized in nanogaps of the microelectrodes and hybridized with their complementary miRNA. Conductance of the enzymatically catalyzed polyaniline nanowires, deposited on hybridized target miRNA, corresponded to the hybridized target concentration. The detection limit of the biosensor as reported by the authors was 5.0fM. Ramanathan et al. demonstrated DNA detection using avidin 33 functionalized polypyrrole nanowires that were electrodeposited in nanometer sized channels on the surface of silicon wafers and the resistance change of the avidin modified polypyrrole nanowires were used as indicator of the binding event of biotin-DNA (Ramanathan et al., 2005). Resistance change for biotin-DNA concentration as low as 1 nM could be detected with an increasing resistance for increasing concentration till 100 nM. 2.3.2 Nanotube Based Biosensors 2.3.2.1 Carbon Nanotubes In recent years, carbon nanotubes (CNT) have found broad applications in biosensors for detecting biomolecules and biological agents. This can be attributed to a number of factors such as unique electrical and electrochemical properties, high mechanical strength, and high aspect ratio for adsorption and immobilization of biomolecules on CNTs. CNTs can be functionalized in different configurations as covalent/non-covalent, defect, sidewall and endohedral, thus, making them compatible for different applications and can also be grown as single-walled nanotubes (SWNT) and multi-walled nanotubes (MWNT) (Figure 2-6) with diameters ranging from 0.4 to 3 nm for SWNT, and from 2 to 100 nm for MWNT respectively (Erickson et al., 2008;Kim et al., 2007b;Vaseashta and mova—Malinovska, 2005). CNTs have two distinct advantages over Si nanowires in their higher electron mobilities and their diameters in sub nanometer range and were first used in FET based devices for chemical sensing by Kong et al. (Kong et al. , 2000). 34 Figure 2-6. TEM images of (A) single-walled carbon nanotubes and (B) multi-walled carbon nanotubes. A SWNT based FET device was developed by Villamizar et al. for the detection of the foodbome pathogen Salmonella Infantis (V illarnizar et al., 2008). The device used anti-Salmonella antibodies adsorbed onto the CNTs as the biological sensing element and was able to detect 100 CFU/ml of the pathogen in 1h. The device was selective and showed no cross-reactivity with antigens such as S. pyogenes and S. sonnei. So et al. reported an aptamer (So et al., 2008) based SWNT FET array as a screening tool for E. coli DHa5 using the most probable number method. The sensor applied RNA based E. coli aptamers as the molecular recognition element and was able to complete the detection process in 20 min. Zhou et al. had utilized the high polarizability and dielectric mobility of SWNTs to capture, concentrate, and detect low numbers of bacteria in small sample volumes using microelectrode arrays (Zhou et al., 2006). AC impedance spectra measurements yielded a detection threshold of 104 bacteria/ml with this technique. SWNT based FET devices functionalized with PNA were also used to detect an RNA sequence of the Hepatitis C virus at concentrations as low as fractional pM range by Dastagir et al. (Dastagir et al., 2007). Dong et al. have shown a label-free detection 35 method for DNA sequences using resistors prepared from double-walled carbon nanotube networks (Dong et al., 2008). The device was able to detect DNA concentrations in the range of 50 and 500 nM and has potential for handheld applications. CNTs have also been extensively applied in the electrochemical detection of biomolecules. CNT based electrochemical biosensors usually employ electrodes modified with CNTS, or hybrid materials, such as CNT-conducting polymer composites, CNT- metal nanoparticles, and other CNT based composites as their detection platforms. Direct electron transfer promoted by the nanotubes between enzymes and the CNT modified electrodes has enabled the widespread application of CNTs in enzyme based electrochemical biosensors (Zhu et' al., 2007). CNT based electrochemical immunosensors and genosensors have also been reported in the literature (Yun et al., 2007;Chang et al., 2008). 2. 3. 2.2 Conducting Polymer Nanotubes Conducting polymer nanotubes are the new class of one-dimensional nanostructures that have found applications as biosensor transducers. Unlike CNT based sensors, conducting polymer nanotube based sensors have the potential of achieving high sensitivity without involving complicated reaction steps, purification or end open processing (Chang et al., 2007b). A polyaniline nanotube array based electrochemical biosensor reported by Chang et al. enabled ultrasensitive detection of nucleic acids (Chang et al., 2007b). The authors demonstrated a sensitivity of 1.0 M target DNA for the biosensor and selectivity of one- nucleotide mismatch at concentrations as low as 137.59 N. Label-free DNA detection using polypyrrole nanotubes has been shown by K0 et al. with a biosensor sensitivity of 36 1.0 nM target DNA (K0 and Jang, 2008). An aptamer conjugated amine-functionalized polypyrrole nanotube based FET device has been developed by Yoon et al. (Yoon et al., 2008) for the detection of the protein, human-a-thrombin. The device could detect 166 nM thrombin in blood serum samples. 2.3.3 Nanoparticle Based Biosm Nanoparticles have unique physical, chemical, Optical, and electronic properties that have made them appropriate for numerous applications in electrochemical sensors and biosensors. Immobilization and labeling of different biomolecules such as enzymes, antibodies, and DNA; acting as electrochemical catalysts and electron transfer agents; and performing the function of reactants and magnetic concentrators are the basic functions performed by these nanoparticles in sensing systems (Luo et al., 2006). Discussed below are current developments in biosensors based on magnetic nanoparticles, metal nanoparticles including gold and silver, and semiconductor nanocrystalline particles (quantum dots). 2.3.3.1 Magnetic Nanoparticles Magnetic nanoparticles (mostly superparamagnetic) have the ability to quickly agglomerate and resuspend in response to changes in external magnetic field. These unique properties of magnetic nanoparticles have led to their exploitation in various separation processes and detection devices such as biosensors. Magnetic separation techniques are widely used in bioengineering and biomedical applications such as magnetic resonance imaging, gene delivery, drug delivery, diagnostics, immunoassays and biosensors (Xu and Sun, 2007;Ludwig et al., 2006;Lee et al., 2006). In biosensor detection, magnetic separation techniques can provide a number of advantages, such as, 37 elimination of nonspecific adsorption of interfering biomolecules by modifying surface chemistries, elimination of sample pretreatment by centrifugation or chromatography thus reducing analysis time, and exclusion of centrifugation steps thus reducing stress-induced damages to biomolecules. Varshney et al. have developed an impedance biosensor based on interdigitated array microelectrode coupled with magnetic nanoparticle antibody conjugates for detection of E. coli 0157: H7 in ground beef samples (Varshney and Li, 2007). The biosensor provided rapid and specific detection of the pathogen with a sensitivity of 7.4 X 104 CFU/ml in pure cultures and 8.0 X 105 CFU/ml in ground beef samples in a detection time of 35 min. Magnetic nanoparticles have also been employed for concentrating and electrochemical sensing of DNA targets in the literature. Zhu et al. have developed a method for detecting DNA hybridization using magnetic nanoparticle modified pyrolitic graphite electrode and the common electrochemical redox couple K3[Fe(CN)6]/K4[Fe(CN)6] (Zhu et al., 2006b). Magnetoswitchable controlled DNA hybridization was performed on DNA probe modified electrodes with subsequent electrochemical detection of the redox indicator. The detection limit for the sensor was 2 nM target DNA. In a separate study, DNA hybridization at magnetic nanoparticles coupled with electrochemical stripping detection of zinc sulfide nanoparticle tags on glassy carbon electrode was used by Zhu et al. (Zhu et al., 2004).The authors were able to attain sensitivity as high as 0.2 pmol /L with the described methods. Lately, several sensing devices have been developed that perform the detection process by measuring the magnetic properties of the magnetic nanoparticle tags. Kaittanis 38 et al. have reported a magnetic nanosensor for detecting Mycobacterium avium sp. paratuberculosis (MAP) in blood and milk samples using magnetic relaxation switches (MRS) (Kaittanis et al., 2007).The detection mechanism involved switching of superpararnagnetic iron oxide nanoparticles between a dispersed and clustered state upon target interaction and measurement of the resultant change in spin-spin relaxation time of the protons in water solution. The authors were able to quantify MAP from 15.5 to 775 CFUs in milk samples using these magnetic nanosensors. A magnetic microarray was developed by Wang et al. utilizing magnetic tunnel junction (MTJ) detectors which was aimed at detecting single molecules of DNA on individual magnetic nanoparticle tags (Wang et al., 2005). Wang and coworkers have also recently prototyped magneto-nano chips using 8 X 8 array of 64 giant magnetoresistance (GMR) spin valve (SV) sensors (Wang and Li, 2008). The authors were also able to successfully detect Human Papillomavirus (HPV) DNA sequences using these GMR DNA chips at concentration levels of 10 pM. A magnetic permeability meter (MPM-100) was developed by Abrahamsson et al for DNA detection (Abrahamsson et al., 2004). The limit of detection of the MPM sensor was 12 ug/ml in buffered solutions. 2. 3.3.2 Gold Nanoparticles In the last few years, intensive research has been carried out in the field of gold (Au) nanoparticle based biosensing devices. Au nanoparticles have a unique physical property of localized surface plasmon resonance i.e. the change in color that arises in the nanoparticles during aggregation (red-to-purple) or redispersion of aggregates (purple-to- red) from interparticle plasmon coupling. The above principle has been applied in Au nanoparticle based colorimetric biosensors for detection of different targets including 39 nucleic acids, proteins, saccharides, metal ions, small molecules and cells. A detailed review on colorimetric biosensing assays has been provided by Zhao et al. (Zhao et al., 2008).The ability of Au nanoparticles to provide a suitable microenvironment for the immobilization of biomolecules retaining their bioactivity is a major advantage in biosensor fabrication. In addition, Au nanoparticles facilitate direct electron transfer between immobilized biomolecules and electrode surfaces without the use of electron transfer mediators, thus making them appropriate for electrochemical biosensing (Pingarron et al., 2008). Panda et al. have developed an optical detection method for the rapid quantification of bacterial cells using fluorescent Au nanoparticle-polymer composites (Panda et al., 2008). The method is based on the loss of fluorescence intensity of positively charged Au nanoparticle-polythiophene composites in the presence of bacterial cells. The detection limit of the system was 1000 bacterial cells for both Gram-positive and Gram-negative bacteria. A highly sensitive electrochemical immunosensor for the detection of S typhi has been reported by Dungchai et al. using copper enhanced gold nanoparticles (Dungchai et al., 2008). The presence of bacteria was detected by measuring the concentration of released copper ions from copper deposited on the colloidal Au nanotags using a copper enhancer solution by anodic stripping voltammetry. The authors reported a detection limit of 98.9 CFU/ml using the electrochemical immunosensor. Wang et al. have described a Quartz Crystal Microbalance (QCM) biosensor based on Au nanoparticles for the real-time detection of E. coli 0157: H7 DNA in a circulating flow system (Wang et al. , 2007c). The sensor was able to detect the target DNA corresponding 40 to 2.0 X 103 CFU/ml of E. coli 0157: H7 cells based on mass change and concomitant frequency shifts of the QCM. Au nanoparticle based biobarcode assays allow highly sensitive detection of DNA sequences and are emerging as promising alternatives to polymerase chain reaction (PCR) for DNA detection. Au nanoparticles heavily functionalized with oligonucleotide probes are the building blocks of biobarcode assays. Hill et al. have reported a biobarcode assay for the detection of B. subtilis genomic DNA (Hill et al., 2007). The biobarcodes were detected by scanometric methods. The assay demonstrated a sensitivity of 2.5 fM for double stranded genomic DNA. A fluorescent bio barcode assay for the detection of the foodbome pathogen Salmonella Enteritidis has been reported by Zhang et al. (Zhang et al., 2009a). Au nanoparticles were first coated with target specific DNA probes and fluorescent labeled DNA barcodes. The DNA target was then sandwiched between the Au nanoparticles and magnetic nanoparticle coated with second DNA probes and separated by applying a magnetic field. The barcode DNA was released from the Au nanoparticles and quantified by fluorescence measurements. The authors reported a sensitivity of 1 ng/ml with the assay. 41 Magnetic nanoparticles with 2 "‘1 probe t” % , W —> ——-> Target DNA Barcode DNA with 1 3‘ probe ——> ——+ Probe Barcode DNA Separation release Figure 2-7. Schematic representation of Au nanoparticle based bio barcode assay (Zhang et al., 2009a). 2.3.3.3 Semiconductor Nanoparticles (Quantum Dots) Quantum dots (QDS) are luminescent nanocrystalline semiconductor particles that are roughly spherical in nature with particle diameters ranging from 1 to 12 nm. QDS have higher fluorescence quantum yields, better photoluminescence stability, and less toxicity than organic fluorophores along with their size tunable fluorescent properties. QDs are also resistant to photobleaching, denaturation of biomolecules and to variations in pH and temperature. Multi color QD nanocrystals have the potential for multiplexed and high-throughput analysis. These properties have made QDS suitable for application in optical detection systems (Costa-Femandez, 2006;Sapsford et al., 2006;Pileni, 2001). QB based optical biosensors can be roughly divided into luminescence-based and absorption- 42 based sensors depending on the changes of emission and absorption intensity (Comparelli etal,2007) Recently, Liu et al. have used multi-layers of quantum dots for rapid magnetic bead based detection of E. coli 0157: H7 eae target DNA (Liu et al., 2008). The target DNA was first captured on magnetic beads by probe hybridization and the signal was amplified by using multi-layers of QDs. Fluorescence intensity measurements yielded a sensitivity of 250 zM within 40-60 min using this method. Su et al. reported an antibody based rapid, sensitive, and specific detection method for E. coli 0157: H7 using cadmium selenide- zinc sulfide (CdSe-ZnS) QDS (Su and Li, 2004b). Magnetic beads coated with anti- E. coli 0157: H7 antibodies were used to capture the target bacteria and were labeled with QDs by streptavidin-biotin conjugation. Fluorescence intensity measurements were proportional to E. coli 0157: H7 concentrations in the range of 103 to 107 CFU/ml with the total detection time being less than 2h. The detection system was also selective against E. coli K12 and S. Typhimurium. Goldman et al. have utilized CdSe—ZnS QDs conjugated with antibodies to perform multiplex fluoroimmunoassays for the detection of cholera toxin, ricin, shiga-like toxin 1 and Staphylococcal enterotoxin B (Goldman et al. , 2004). QDs have also found suitable applications in fluorescence resonance energy transfer (FRET) biosensors. The FRET technique relies on distance-dependant transfer of energy from a donor fluorophore to an acceptor fluorophore. The narrow emission spectra and the broad absorption spectra of QDS combined with their donor-acceptor separation distance of 1-10 nm result in improved performance of QDS in FRET assays over organic fluorophores. Zhang et al. reported an ultrasensitive DNA nanosensor based on FRET 43 using single QDs (Zhang et al., 2005a). The system used QDs conjugated with DNA probes to capture DNA targets bound to dye-labelled reporter strands which formed a donor-acceptor ensemble. The QDS acted as a target concentrator and signal amplifier in the nanoscale domain. The nanosensor was capable of generating a strong FRET signal when bound to 50 copies of DNA or less. Recently, a FRET biosensor based on Au nanoparticles and QDS was developed by Stringer et al. for the detection of Porcine Reproductive and Respiratory Syndrome Virus (PRRSV) (Stringer et al., 2008). The biosensor consisted of two architectures; a PRRSV- specific capture antibody labeled with a fluorescent dye bound to a QD labeled protein and a PRRSV- specific capture antibody labeled with a fluorescent dye bound to a Au nanoparticle labeled protein. The detection limit of the sensor was 3 particles /ul. The metal components of the semiconductor nanoparticles in the QDS (CdS, PbS, and ZnS) have intrinsic redox properties susceptible to sensitive electrochemical stripping analysis thus making them suited for electrochemical biosensors. QD based electrochemical biosensors have been reported in literature for detection of different protein molecules and DNA but have not been applied in the detection of pathogenic microorganisms (Thurer et al., 2007;Hansen et al., 2006). Pumera et al. have developed a protocol for the detection of DNA hybridization based on magnetically triggered electrochemical detection of Au67 QD tracers (Pumera et al., 2005). The authors reported a detection limit of 12nM target DNA with this system. 2. 3.3.4 Silica Nanoparticles Literature reveals that silica (Si) nanoparticles have been successfully used as nanotransducers in electrochemical and fluorescence based biosensors and bioassays. 44 Zhao et al. have reported a rapid bioassay for the detection of a single bacterial cell of E. coli 0157: H7 using fluorescent-bioconjugated silica nanoparticles (Zhao et al., 2004). The authors were able to detect a single E. coli 0157: H7 bacterium within 20 min in spiked ground beef samples with this bioassay. Multicolored FRET Si nanoparticles were used by Wang et al. for multiplex monitoring of bacterial pathogens (Wang et al., 2007b). Each dye doped Si nanoparticle encapsulated about 10,000 dye molecules in a 60 nm silica sphere which provided an extremely strong fluorescent signal for bioanalysis. The authors reported simultaneous detection of E. coli, S. Typhimurium and S. aureus with this assay in 30 min. Ping et al. have developed a PCR based detection system for the detection of a SARS Coronavirus gene using Si coated superpararnagnetic nanoparticles and Si coated fluorescent nanoparticles (Gong et al., 2008). PCR amplified targets were concentrated using Si coated magnetic nanoparticles and then hybridized with the Si coated fluorescent nanoparticles in a sandwiched hybridization format. The limit of detection of the assay was 2.0 X 103 copies in a total detection time of less than 6h. A sensitive electrogenerated chemiluminescence (ECL) detection of DNA hybridization based on Ru(bpy)32+ doped Si nanoparticles was developed by Chang et al. (Chang et al., 2006). The Ru(bpy)32+ doped Si nanoparticles labeled DNA probes were first hybridized with DNA targets immobilized on the surface of polypyrrole modified platinum electrodes. The hybridization events were then detected by ECL signals. Using the ECL signals, the authors were able to detect the target DNA at levels as low as 1.0 X 10'13 mol/l with selectivity of a three base mismatch sequence and a non-complementary sequence. 45 2.3.4 Nanoporous Material Based Biosensors Novel nanoporous materials such as porous silicon, porous carbon, and porous alumina have recently found effective applications as biosensor transducers. Nanoporous silicon, characterized by its high surface to volume ratio and photoluminescence properties, has been employed in more than one biosensor transduction mechanisms. Mathew and Alocilja developed a pathogen biosensor for the detection of E. coli exploiting the optical properties of nanoporous silicon (Mathew and Alocilja, 2005). The reaction of B-galactosidase enzyme from E. coli with a dioxetane-polymyxinB mixture functionalized on porous silicon chips generated light at 530 nm. Light emission measurements from the porous silicon biosensor demonstrated a sensitivity of 101-102 CFU/ml for the pathogen. Electrochemical detection of targets using nanoporous silicon platforms has been reported by Zhang et al. (Zhang and Alocilja, 2008). The authors fabricated nanoporous silicon with P-type silicon wafers by electrochemical etching and applied them in the detection of DNA targets for the foodbome pathogen Salmonella Enteritidis. The sensitivity of the biosensor based on the electrical properties of DNA, redox indicators, and cyclic voltammetry was 1 ng/ml. The extremely high surface area of nanoporous silicon films have also been used for sensitive detection of the bacteriophage virus MSZ by Rossi et al. (Rossi et al., 2007). Porous silicon films were used to selectively capture dye-labeled M82 viruses from solution and a sensitivity of 2 X 107 plaque-forming units per ml (PFU/ml) were reported by the authors by measuring fluorescence from the exposed porous silicon film. Bothara et al. (Bothara et al., 2008) have developed an electrical immunoassay based on nanoporous alumina membranes for the sensitive detection of proteins. The detection 46 principle involves the formation of electrical double layer and the perturbations caused by proteins trapped in a nanoporous alumina membrane over a microelectrode array platform. Label-free electrical detection of protein biomarkers for cardiovascular diseases, C-reactive protein and (CRP) and myeloperoxidase (MPO) were achieved through this assay. The lower detection limits for CRP and MP0 were 200 pg/ml and 500 pg/ml respectively. Porous conductive carbon has been shown to be a good matrix for biosensor construction. The high conductivity of carbon is suitable for electrochemical transduction and in combination with its porosity makes it ideal for adsorption Of biomolecules. Nanoporous carbon has been successfully used for the construction of acetylcholine esterase biosensor by Sotiropoulou and Chaniotakis (Sotiropoulou et al. , 2005). 2.3.5 Nanorod and Nanosghere Based Biosensors Gold nanorods have been successfully employed in optical biosensors. Marinakos et al. have developed a label-free chip-based localized surface plasmon resonance (LSPR) biosensor using gold nanorods (Marinakos et al., 2007). The biosensor monitored streptavidin binding to biotin by studying the LSPR peak shifts at the gold nanorod surface. The biosensor sensitivity was 94 pM in phosphate buffered saline and 19 nM in serum. Multiplex detection of three different IgG molecules was also achieved on gold nanorods of different aspect ratios fabricated by a seed-mediated growth by Yu and Irudayaraj (Yu and Irudayaraj, 2007). A gold/silicon hetero nanorod biosensor was developed by Fu et al. for detection of the foodbome pathogen Salmonella Typhmurium (Fu et al., 2008). The hetero nanorods were functionalized with anti-Salmonella 47 antibodies and detection was accomplished through fluorescence signaling mediated by multiple Alexa488 dye molecules bound on the nanorods. Different nanospheres viz. silica, nickel oxide (NiO), cadmium sulfide (CdS), and magnetic gold are being currently utilized as bioSensor transducers in literature. NiO hollow nanospheres have been used for glucose biosensing by Li et al. (Li et al. , 2008). A novel nitrite biosensor based on direct electron transfer of hemoglobin (Hb) immobilized on CdS hollow nanosphere modified glassy carbon electrode has been reported by Dai et al. (Dai et al., 2008). Thionine doped magnetic gold nanospheres synthesized by reverse micelle method have enabled ultrasensitive electrochemical immunosensing of carcinoembroyonic antigen protein by Tang et al. (Tang et al., 2008). The use of fluorescent silica nanospheres as luminescent signal amplifiers in detecting the breast cancer marker HER2/neu in a simple glass slide based assay has been reported by Rossi et al. (Rossi et al. , 2006). 48 2.4 POLYANILINE — A CONDUCTING POLYMER 2.4.] Introduction Polymers containing loosely held electrons in their backbones are referred to as conjugated polymers or conducting polymers and have attracted considerable interest for the past two decades. Research on conducting polymers intensified after the rediscovery of polyacetylene in 1977 by McDiarmid and Heeger. They were able to enhance the electrical conductivity of polyacetylene by many orders of magnitude by simple doping with oxidizing and reducing agents (Gerard et al., 2002). The presences of u—electron backbones in the conducting polymers result in their unusual electronic properties, such as high electrical conductivity, low ionization potential, and high electron affinity. They are also known as ‘intrinsically conducting polymers’ due to the presence of partially filled molecular orbitals which allow free movement of electrons throughout the polymer lattice and overlapping molecular orbitals which allow the formation of delocalized molecular wave functions. These types of polymers have also attracted interest for their nonlinear Optical properties, since the electrons in these delocalized systems are easily polarized by an external electric field. These polymers have been considered to be as “molecular wires” for nanotechnology (German and Grubbs, 1991). As a result of such unusual properties, conducting polymers have a number of practical applications, such as in electronic displays, telecommunication, electrochemical storing systems, and molecular electronics and sensors (Hohnholz et al., 2005;Kaneto et al., 2007;Ramanavicius et al., 2006). Table 2-1 shows the chemical structures and conductivity of some common conducting polymers. 49 Table 2-1. Chemical structure and conductivity of typical conducting polymers (adapted and modified from Rahman et al., 2008). Conducting Polymer Structure Conductivity (S/cm) Polyacetylene N ~ 1000 n Polyparaphenylene ‘83} 100 ~ 500 n | Polyparaphenylene ~3 vinylene n Polyazulene ~0 .l O n ,- l‘ Polyaniline --©—N 1~ 100 x n Polypyrrole N 40 ~ 100 Polythiophene I :3: ‘ 10 ~ 100 n Polycarbazole O N O 10 ~ 100 .1. n NH2 Polydiaminonapthalene 10 '3 NH2 50 2.4.2 Structure and Electrical Progerties Among all conducting polymers, polyaniline is intensely investigated particularly due to its excellent stability in different solutions, good electronic properties, and strong biomolecular interactions (Feast et al., 1996;Ryder et al., 1997). Polyaniline is synthesized from its monomer form, aniline. There are two basic methods of polyaniline synthesis: chemical synthesis and electrochemical synthesis. The chemical synthesis involves step-growth polymerization of an aqueous solution of the monomer in the presence of an oxidizing agent and a protonating acid. In the aqueous acidic media, aniline cation radicals are the first product of the oxidation. The aniline cation radicals then recombine into benzidine and N-phenyl-p—phenylenediamine, or, participate in the growth of polyaniline chains in the perrrigraniline form. The oxidative polymerization reaction continues according to equations 2-3 and 2-4 until all the pemigraniline is converted into emeraldine by equations 2-5 and 2-6 (Stej skal and Gilbert, 2002). .NH; ‘__—-_, .NH2+ H++ e (2-3) R‘NHz + .NHZ ‘__——:_—; +. +. + RHNNH2 + 2H + 26 (2-4) 51 820:} 26 3:; 2804' (2-5) tear-t +e -——» Ml (2-6) The electrochemical synthesis of polyaniline is carried out in a typical electrochemical bath by adopting a standard three-electrode configuration. The standard three-electrode system usually comprises of a working electrode, a counter electrode and the reference electrode. The electrochemical polymerization generally employs a constant current or galvanostatic, constant potential or potentiostatic or potential scanning/cycling methods (Ahuja et al., 2007). The commonly used working electrodes are chromium, nickel, gold, platinum, palladium and glass coated with tin oxide or indium tin oxide. 2. 4. 2.1 Oxidation States The conducting polymer, polyaniline, exists in a variety of forms differing in electrical conductivity and color. The general chemical structure of polyaniline has been H O 1* C I N shown in Figure 6, where, indicate the reduced repeat units, and -~-©=~O indicate the oxidized repeat units of the polymer. 52 H H N N = N Reduced unit Oxidixed unit 14y — ‘ K Figure 2-8. Structure of polyaniline, x = degree of polymerization, [y + (l-y)] = 1. (Ray et al., 1989) There are three distinct oxidation states of the polymer, namely, “leucoemeraldine”, “emeraldine” and “pemigraniline”. The fully reduced form of polyaniline is “leucoemeraldine” (l-y = 0, y =1), while “pemigraniline” (l-y = 1, y = 0) is the fully oxidized form of polyaniline and “emeraldine” (y = l-y = 0.5) is the 50% oxidized form of polyaniline (Ray et al. , 1989). Each oxidation state of the polymer can exist in its base form or its protonated form (salt) by treatment of the base with acid. Interconversion between the different forms of polyaniline occurs during the oxidative polymerization of aniline which imparts electrical conductivity and color to the different forms. The most important and stable form of polyaniline is protonated emeraldine which is green in color and electrically conductive. It is produced directly by the oxidative polymerization of aniline. When stronger oxidizing conditions are employed, it converts to protonated pemigraniline which is blue and is expected to be conducting. Further treatment with alkali results in pemigraniline base which is violet and non-conducting. The acid-base transition occurs at pH 0-1. Emeraldine can also be reduced to the colorless, non-conducting, leucoemeraldine form (Stej skal et al., 1996). 53 2. 4. 2.2 Electronic Conduction It is well known that the electrical conduction properties of elemental semiconductors, such as Si, can be rigorously controlled by addition of small quantities of foreign atoms into the semiconductor lattice. The semiconductors can be made n type or p type depending on the nature of the dopant atoms, i.e. whether it has an excess or deficit of electrons. The electrical conduction in conjugated or conducting polymers can also be explained by a similar terminology. However, the doping mechanism in conducting polymers is considerably different from the conventional semiconductors. Firstly, the doping levels are significant in conjugated polymers (as large as 10 mole percent) and secondly, there is charge transfer between the incorporated dopant atom and the polymer chain, hence the later is partially oxidized or reduced (Lyons, 1994). The partial oxidation of the polymer chain is termed as p-doping while partial reduction is termed as n-doping. Electrical conductivity results (0) from the existence of charge carriers and the ability of those carriers to move. It is expressed as, c = nep (2-7) where, n is the number of charge carriers per unit volume, e is the electronic charge and u is the carrier mobililty. Hence, doped conjugated polymers or conducting polymers are good conductors for two reasons; firstly, doping introduces charge carriers into the 112- electron system of the polymer and since every monomer is a potential redox site, they can be heavily doped, and secondly, the 712- bonding leads to 7t- electron delocalization along the polymer chains thereby leading to charge carrier mobility which is extended 54 into three dimensional transport by interchain electron transfer interactions (Heeger and Smith, 1991). The half-oxidized form of polyaniline (i.e. emeraldine) is the most studied and the most stable form of polyaniline. The emeraldine base form of polyaniline has been found to be different from other polymers in several aspects. First, it is not charge conjugation symmetric in contrast to polymers such as polyacetylene and polythiophene. Second, both the carbon rings and the nitrogen atoms in polyaniline are within the conjugation path forming a generalized “A-B” polymer. Third, the C6 benzenoid rings of polyaniline can rotate or flip thus significantly altering the electronic structure (Pouget et al. , 1991). It has been shown that the emeraldine base form of polyaniline can be reversibly varied from an insulator (o < 10'10 ohm'lcm'l) to a conductor (0 ~ 10 Ohm’lcm'l) through protonation (Epstein et al., 1987). The doping of emeraldine base with a non-oxidizing protonic acid such as HCl involves a new concept. Unlike the doping of all other conducting polymers, the number of electrons in the polymer backbone remains unchanged (Macdiarmid et al. , 1987). Unlike other doped conducting polymers which are polycarbonitun ions or polycarbanions, the emeraldine salt polymer is based on a “nitrogenonium” ion polymer in which the positive charge resides chiefly on the nitrogen (Macdiarmid et al. , 1987). The conductivity of polyaniline is believed to be due to the analogous protonation of emeraldine base which leads to the formation of a polysemiquinone radical cation (polaron) and gives rise to a polaron conduction band by coulombic repulsion leading to a metallic state (Macdiarmid et al., 1987). 55 i“ N/ l j, H Protonation (b) 1' <2x>ll+ T N i ii i“ l H H ' x (c) Internal Redox Reaction H H | I N ”/31. ll 1 l H H X Polaron Separation (d) Ill H | ‘1‘!- ”A"! \i i J H H 2X Figure 2-9. Structure of emeraldine (polyaniline), (a) before protonation and (b) - (d) after protonation, (b) formation of bipolarons, (c) formation of polarons, and (d) the separate polarons which result in a polaron lattice (adapted from Stafstrom et al., 1987). 56 A model based on two-step transition from isolated, doubly charged spinless, bipolarons to a polaronic metal has been suggested by Stafstrom et al (Stafstrom et al., 1987). The first step relates to instability of a bipolaron on a polyemeraldine chain which results in the formation of two polarons, and the second step involves the separation of the polarons to yield a polaron lattice (Stafstrom et al., 1987). Figure 2-9 shows the changes in the structure of polyaniline as suggested in the above model. Figure 2-10 shows the band structure of a conjugated polymer as a function of doping level which illustrates the polaronic and bipolaronic states in the band gap. A neutral polymer acts as an insulator and has a full valence band and an empty conduction band which is separated by a band gap. Removal of an electron by doping results in the generation of a polaron level at the midgap. Further oxidation results in the removal of a second electron to generate a bipolaron. CONDUCTION BAND I E9 A VALENCE BAND | I Neutral polymer Polaron Bipolaron Via 1e- oxidation Via 2e-oxidation ‘— Doped polymer _’ Figure 2-10. Band structure of conjugated polymers showing polaronic and bipolaronic states. 57 2.4.3 Electrochemistgy and Redox Switching The electrochemical preparation of polyaniline has typically been carried out in aqueous electrolytes galavanostatically, potentiostatically or through cycling of the potential of the substrate anode. Several electrolytes have been successfully used for the synthesis of polyaniline such as hydrochloric acid, perchloric acid, sulfuric acid, nitric acid, trifluoroacetic acid and fluoroboric acid (Delvaux et al., 2000;Duic et al., 1994;Duic et al., 1995;Pan et al., 2006;Wei and Ivaska, 2006)with the polymer being deposited on diverse substrates such as platinum, gold, indium tin oxide, carbon and stainless steel (Bereket et al., 2005;Mazur et al., 2002;Tang et al., 1996;Wei et al., 2008;Zhang et al., 2009b). Investigation of the electrochemical properties (redox switching) of polyaniline is considered to be the key factor in understanding the physical and chemical properties of polyaniline. The redox switching process of polyaniline has been studied for many years using various kinds of voltammetry. The concomitant processes that occur during the redox switching of polyaniline are as follows: (a) cyclic voltammetry reveals that two peaks are associated with the switching of polyaniline from the fully reduced leucoemeraldine state to the conducting emeraldine state and from the emeraldine state to the fully oxidized pemigraniline state; (b) the conducting emeraldine state of polyaniline exits in a potential window instead of a fixed potential; (c) the redox switching property of polyaniline depends mainly on the pH of the medium and in neutral or alkaline medium, the polymer loses its electrochemical activity; ((1) the redox behaviour of polyaniline is fundamentally asymmetric; and (e) the oxidation transition occurs at a lower rate than that of the 58 reduction transition (Gospodinova et al., 1996;Grzeszczuk and Szostak, 2003;Hong and Park, 2005). The redox transition of polyaniline usually occurs in the potential range of -200 to 400mV (saturated calomel electrode) and is accompanied by proton ejection or injection. There is a proton ejection during the oxidation of a fully relaxed polymer from the leucoemeraldine form to emeraldine form and from it to fully oxidized form, while the reduction of the polymer is accompanied by protOn injection which is incomplete because of the slow proton equilibration process (Ybarra et al., 2000;Lyons, 1994). The redox switching in polyaniline is also accompanied by a small increase in volume of the polymer. Several factors such as water and ion exchange with the electrolyte, coulombic repulsion between charged sites in the polymer backbone, anion-polymer interactions, and a structural change of the polymer backbone govern the swelling of the polymer during the switching process (Lizarraga et al., 2004). The effect of pH on the electrochemical activity of polyaniline has been studied by authors (Prakash, 2002). Cyclic voltammograms (CVS) of polyaniline films on a Pt working electrode and Ag/AgCl reference electrode were recorded in HCl solutions in the pH range of 1-4 by the author (Figure 2-11). At pH 1.0, three reversible anodic peaks and their corresponding reduction peaks, were observed. As the pH was increased, a shift in one redox peak toward lower potential values which merged into a single broad peak at pH 4.0 occurred This peak was associated with the involvement of protons in the redox reaction which made the peak pH dependant, whereas, the other peaks only involved transfer of electrons (Prakash, 2002). 59 A83 .iéemv a... 3 Ea 3. A3 .3 3 .3 3V .3 3 a as? an a 828.... as... «5:522. a £6 .:-~ 2%: Em Em . ad. ad ad ad ad Nd ad. ad ad dd dd N; u u u . Q.°I - q u u - V.°I I ad. 4 Nd- - 3 I I o... I .. dd W I Nd W I ad I vd o I dd u I ad ad ad Em Em Em nd. ad ad wd md N... ad. ad ad md dd N... Nd- dd ad ad ad N... u q u u . ¢.°I A u u u u 0.0: u u u d u 9.? I Nd. I ed. #6. - o... . N5. 1 N5. . N... n . o... ) . n . w m / a o w l #0 i. J N.° N.° I. o I M.” a I vd ed o o od 60 2. 4.3.1 Cyclic Voltammetry Cyclic voltammetry is a form of linear sweep voltammetry in electrochemistry where the potential applied to the working electrode of a three electrode electrochemical cell is ramped linearly versus time between an initial potential (E 1) and a final potential (E2). When the potential reaches E 2, it is ramped back to the initial potential of E]. This ramping of the potential is known as the scan rate and is usually expressed in V/s or mV/s. For a reversible redox system, the resulting current (I) measured while scanning the potential is represented with respect to the potential (E) in a typical cyclic voltammogram. Figure 2-12 shows a typical I-E curve (cyclic voltammogram) for a redox system. The voltammogram is characterized by a peak potential Ep, a potential corresponding to the point where the measured current reaches its maximum value Ip. In Figure 2-12, [pa and 1pc indicate the anodic and cathodic peak currents and End and Epc indicate the anodic and cathodic peak potentials, respectively. Two parameters that are of interest in the I -E curves are the ratio of the peak currents [pa/1pc, and the separation peak potentials, Epa —Epc. For a reversible system, the peak current, IP, is given by the equation 1/2 F3 . 1p = 0.4463[fi] n3/2AD3/2C0V1/2 (2-8) where, F = Faraday’s constant (Q mol'l), R = universal gas constant (J mol'lK'l), T = temperature (K), n = no. Of electrons exchanged in the reaction, A = area of the electrode surface (cmz), D0 = diffusion coefficient of the electroactive species (cmz/s), C5: 61 concentration of the electroactive species (mol/cm3), and v = scan rate (V/s) (Bard and Faulkner, 2001). Similarly, the peak potential, Ep, for a reversible system is given by the equation RT Ep =E,,2—1.109E (2-9) For systems with stable product (reversible reactions), the following two conditions should be satisfied: the anodic peak current should be equal to the cathodic peak current, i.e., [pa/1pc = 1 and the separation of the anodic and cathodic peak potentials should obey the following equation: IAEPI = lEpa—Epc = 0.059/n (2-10) For totally irreversible redox systems, Ep is a function of scan rate, shifting (for reduction) in a negative direction by an amount of about 0.03V ten fold increase in v whereas in quasi-reversible systems, 1,, is not proportional to v (Bard and Faulkner, 2001). 62 I (A; Cathode so?) Anode Figure 2-12. Typical cyclic voltammogram (I-E curve) for a redox system. Cyclic voltarnmtery is a very powerful tool for electrochemical studies of new systems and provides useful information on complicated electrode reactions. The current research will investigate the diagnostic strength of cyclic voltammetry in evaluating the concentration of an electrically active species from peak current heights. 2.4.4 ElectricgflLv-Active Magnetic Polyaniline Magnetic polymer nanostructures are a new class of materials where magnetic nanoparticles are embedded in polymer matrices. Such nanostructures have attracted considerable interest in recent years due to their immense potential for applications in electromagnetic devices, drug targeting, cell separation, enzyme immunoassays and electromagnetic interface shielding (Asmatulu et al., 2005;Boissiere et al., 2006;El- Tantawy et al., 2004). Multi component systems containing conducting polymers and nanoparticles of metal oxides have the advantages of non-corrosiveness, light weight, 63 mechanical strength and dielectric tunability along with novel magnetic and optical properties (Poddar et al., 2004b). Polyaniline is the most studied conducting polymer in such multi component systems due to its unique properties such as controllable electrical and chemical properties, excellent environmental stability and simple and low cost synthesis (Ray et al., 1989). The general approach for the synthesis of such magnetic polyaniline nanocomposites or nanoparticles is polymerization of the monomer aniline around a magnetic core or template nanoparticle as represented in Figure 2-13. Oxidizing agent ””2 Acid Template Aniline Magnetic polyaniline Figure 2-13. Synthesis of electrically-active magnetic polyaniline (Li et al., 2007b). Till date, the most common magnetic nanomaterials that are used in the synthesis of magnetic polymer composites are iron oxides and ferrofluids. A review of the magnetic properties of iron oxides has been provided by Kryszewski et al (Kryszewski and Jeszka, 1998). Some of the most important properties as mentioned by them are: (a) electrostatic exchange dominates for Fe ions at adjacent sites, (b) the specific saturation magnetization of iron oxide nanoparticles decrease with increase in specific surface of the nanoparticle and is also influenced by the particle morphology, (0) surface hydroxyl groups by adsorption of water make iron oxide nanoparticles amphoteric so that they can react both with acids and bases, and (d) the acidity and reactivity of differently coordinated iron 64 changes. According to Kryszewski et al. the incorporation of iron oxide nanoparticles into polymer matrices involves specific interactions with ligands which replace the surface hydroxyl groups as represented in Figure 2-14. L L L | / \ / 777- 0 777— 770777—077 770777077- | \ Fe / I | Fe Fe Fe Mononuclear Mononuclear Binuclear Figure 2-14. Polymer ligand interactions to iron oxide particle surfaces (L = ligand) (Kryszewski and Jeszka, 1998). Under ideal circumstances, nanoparticles can exhibit novel magnetic properties when they exist as isolated particles. A critical obstacle in maintaining a nanoscale magnetic material is its tendency to agglomerate (Kryszewski and Jeszka, 1998). For practical applications, a collection or aggregate of nanoparticles or nanopowders are more prevalent. The dispersion of the nanoparticles in the polymer matrix is also a critical issue and needs extensive research. The competition between polymer-polymer and polymer- particle needs to be balanced in order to achieve good dispersion and avoid clustering of particles in the polymer composites (Alam et al., 2007). Recent literatures suggest that different kinds of magnetic nanoparticles are being used as a core or template for the synthesis of electrically active magnetic polyaniline nanostructures. Table 2-2 shows a brief review of electrically active magnetic polyaniline nanostructures available in the literature. The size range of the synthesized nanostructures varies from 30 nm to 5 pm and their magnetization values are in between 0.76 and 181 .I emug . 65 Guiana: Eimmfiom n IO¥ 62..ch :5 no: n :08 .28 2822533889 u «.me .28 ozoaamoi n 5.3a .22. 2528.3: u 6: .28 28:3 222292 n .4sz 932 .5 as set S: 2 .06.; mom ABS :3 3 been: 2.33 1 Oman seam Aegean mfiomaaafse as a accuse 3an aaaam 6m SEN dad Ea gee 31: YN 52% .2 sac 406.585 an seem as a wee 33.2 832 ”28 .2 “an “408%va BEES MOE seem as a med: 2 8-2 dam doom u“sou dds._flaggeezmdaecoaz: 6: Goom re a E 338 3.8 cowomSeanzv sEomaNz 6m seem as a see 35-on z and engages: 8: .4me 388 as a weed 92-3 83% ESE 3-9. a5 406m scam-» 6.1m seem as a Bee 98-9% 2 comments aces a: 8: 2538. A88 :3 a mace 22%? 872 A406”: sea a: .5 no: 6: 388 as a E 4.9: 885% EOE 8: 596% $2. a codenamed SEANHHMNE A55 aim—SEE Sidney—EEO Eamon— .moEuonmcnan 03:5“on ate—as:— o>=oaJm=ao€uoflo he 32.5% .NIN 035—. 66 2.4.5 Polyaniline in Biosensor Apglications Conducting polymers such as polyaniline have gained. popularity in biosensor applications due to the following reasons: 0 Flexible chemical structure that allows modulation of the required electronic and mechanical properties of the polymer. 0 Suitable immobilization matrix for biomolecules providing suitable environment for the immobilization. 0 Associated with efficient transfer of electric charge produced by biochemical reaction to electronic circuits. 0 Can be deposited onto electrode surfaces which provide them with the versatility of being used in diverse biosensor platforms. 0 Control over parameters such as polymer layer thickness, electrical properties and bio-reagent loading. Langer et al. have developed an enzyme based biosensor for detection of the amino alcohol, choline, using nanostructured polyaniline layers of controlled porosity and micrometer or nanometer thickness (Langer et al., 2004). The enzyme choline oxidase was immobilized in the nanoporous polyaniline layers which were useful in trapping the enzymes due to coulombic interactions between the positively charged polyaniline and negatively charged polar groups in the enzyme molecules. The authors were able to enhance the electrical response of the biosensor by continuous charging and discharging of the polyaniline molecules in presence of the enzyme. The sensitivity of the sensor as reported by the authors was 5 pA mmol/L in the amperometric mode and 10 mV mmol/L 67 in the potentiometric mode, with the detection limit being 20 mmol/L for choline in the potentiometric mode. An antibody based polyaniline biosensor was developed by Muhammad-Tahir et al. for detection of the bovine viral diarrhea virus using indium tin oxide (ITO) as the platform (Tahir et al., 2007). The ITO glass was spin coated with self-doped polyaniline which served as a matrix for immobilization of antibodies specific to the Virus. The antibody modified electrodes were then treated with 1 ml of the viral culture for 30 min. The detection process involved measurement of the amperometric response of the antibody modified polyaniline spin coated ITO electrodes before and after treatment with , the viral culture in a three electrode electrochemical cell. Malhotra and coworkers have reported polyaniline based nucleic acid sensors for detection of E. coli and Mycobacterium tuberculosis genomic DNA. Avidin modified polyaniline was electrochemically deposited on a Pt disk electrode and immobilized with biotin labeled E. coli probes (Prabhakar et al., 2008;Arora et al., 2007). Direct detection of E. coli was performed using differential pulse voltammetric technique in the presence of methylene blue as DNA hybridization indicator. The authors were able to detect 0.01 ng/uL of E. coli genomic DNA and 11 E. coli cells/mL within 605 to 14 min. Detection of M. tuberculosis genomic DNA was achieved by the same authors using polyaniline modified gold electrodes and PNA as the bio-recognition element. The authors reported a detection limit of 0.125 X 10'18M in a 5 min sonicated M. tuberculosis genomic DNA within 1 min of hybridization time. Polyaniline has been used in numerous biosensing applications, viz, amperometric, potentiometric, conductometric, optical, calorimetric and piezoelectric biosensors. 68 Different biomolecules ranging from enzymes, microorganisms, and antibodies to nucleic acids have been employed in the polyaniline based biosensors. Detailed information on the different polyaniline and conducting polymer based biosensors can be found in review articles by Wei et al., Malhotra et al., and Gerard et al. (Wei and Ivaska, 2006;Malhotra et al., 2006;Gerard et al., 2002). 69 CHAPTER 3: RESEARCH HIGHLIGHTS 3.1 RESEARCH NOVELTY The innovativeness of this research lies in exploring the prospect of a novel nanostructured material (electrically active magnetic polyaniline, EAPM, nanoparticles) in the dual function of a magnetic concentrator as well as a biosensor transducer. The dual potential of this novel nanomaterial was exploited in the development of a nano- biosensor for food safety and biodefense. The versatility of the nanomaterial was demonstrated in both antibody based and DNA based biosensor detection systems. The nanobiosensor described in this study is novel in terms of design and application. Current literature shows a considerable amount of research on polyaniline and magnetic polyaniline synthesis. A sizeable number of applications can be found in the use of polyaniline transducers in biosensors as well. Extensive literature is also available in magnetic concentration procedures using different micro- and nano- magnetic particles. However, till date no literature has been reported that employ EAPM nanoparticles in an integrated design of biosensor transduction and magnetic concentration in one system. Table 3-1 demonstrates the novelty of this research in contrast to existing literature. The shaded regions indicate the absence of any current literature in these areas and the contribution of this research to the knowledge base at this point of time. 70 Table 3-1. Novelty of research as compared to existing literature Parameters Current Comments literature Polyaniline synthesis (Li et al. , One dimensional nanostructures, 2007a) vanadic acid oxidant, size-10/50 nm Polyaniline based antibody (Sai et al. , Piezoelectric immunosensor, IgG biosensor 2006) detection Polyaniline based DNA (Chang et al. , Alumina electrode, electrochemical biosensor 2007a) detection of synthetic Oligos Electrically active magnetic (Li et al. , Polyaniline-hydroxyl iron, diameter: polyaniline (EAPM) synthesis 2007b) 0.5-5 pm EAPM based antibody biosensor EAPM based DNA biosensor EAPM nanoparticles as redox indicators in DNA detection Magnetic particles as cell (Yang and Li, Impedance detection of Salmonella, concentrator in biosensor 2006) Dynabeads®, diameter-2.8 pm Magnetic particles as DNA (Yeung and Electrochemical detection of E. coli, concentrator in biosensor Hsing, 2006) commercial beads, diameter-3 um EAPM as cell concentrator EAPM as DNA concentrator 71 3.2 ESEARCH SIGNIFICANCE Electrically active polyaniline coated magnetic (EAPM) nanoparticles couple the flexibility of the conducting polymer, in terms of its chemical, electrical and mechanical properties, with the magnetic properties of the core material, iron oxide. EAPM nanoparticles when adopted in biosensors are expected to provide several advantages in detection. The nanoscale dimensions of the EAPMS will impart an increased surface to volume ratio for biological events to occur. Appropriate biological surface modification of the EAPMs can bring them in close proximity to targets for achieving fast assay kinetics, e.g., rapid antigen-antibody interactions. In addition, magnetic manipulation Of the EAPMs can minimize matrix interference due to improved separation thus avoiding pre-enrichment and pre-treatment procedures commonly practiced in food and environmental samples. Hence, magnetic manipulation will be advantageous in reducing sample processing time and background signal in detection devices. Furthermore, the electrical, electrochemical, and magnetic properties of the EAPMs will present multiple options in designing biosensor platforms. The compatibility of the polymer surfaces in EAPMS with biomodification procedures will add to the versatility of these biosensor platforms. . As described above, EAPM nanoparticles hold great promise in future biosensing applications. The current research developed an antibody based and a DNA based EAPM biosensor with high sensitivity and specificity and rapid detection time. Two different biosensor platforms involving different transducing mechanisms were examined. 72 3.3 HYPOTHESIS This research was based on the following hypothesis: 3.4 Electrically-active polyaniline coated magnetic (EAPM) nanoparticles can be synthesized and modified with different biological detecting elements such as antibodies and DNA. The magnetic properties of the EAPM nanoparticles will concentrate biological targets such as whole cells and DNA. The electrical properties of the EAPM nanoparticles can be exploited in different biosensor transducing mechanisms to report a biodetection event. The dual function of the EAPM nanoparticles can be demonstrated using B. anthracis as the model pathogen RESEARCH OBJECTIVES The specific Obj ectives of this research are as follows: Objective 1: Synthesis and characterization of EAPM nanoparticles. Objective 2: Design and fabrication of an EAPM based immunosensor for detection of B. anthracis spores. Objective 3: Evaluation of sensitivity and specificity of the EAPM based immunosensor and determination of EAPM immunomagnetic capture efficiency Objective 4: Fabrication of an EAPM based electrochemical DNA biosensor for detection of B. anthracis. Objective 5: DNA hybridization on EAPM NPS and sensitivity and specificity evaluation of the EAPM based electrochemical DNA biosensor. 73 CHAPTER 4: RESEARCH MATERIALS & METHODS 4.1 OBJECTIVE] Synthesis and characterization of electrically active polyaniline coated magnetic (EAPM) nanoparticles (NPS) This objective was aimed at synthesizing the EAPM NPS following a chemical synthesis procedure and performing magnetic, electrical, and structural characterization studies of the NPs for assessing their suitability and applicability prior to developing the biosensors. 4.1.1 EAPM Nanogarticle Synthesis The EAPM NPS were synthesized from aniline monomer made electrically active by acid doping and gamma iron (III) oxide (y-Fe203) NPS (Sharma et al., 2005). The 7- FezO3 NPS were obtained from a commercial source (Sigma-Aldrich, MO). These NPS were dispersed in a mixture of 50 ml 1M HCl, 10 ml de-ionized water and 0.4m] of aniline by sonication for 1h at 0°C to disintegrate the agglomerated nanoparticles. This was followed by a slow drop-wise addition of the oxidant, ammonium persulfate, at a rate of 0.1 ml/min with continuous stirring of the above solution mixture which resulted in coating of polyaniline on the smaller y-Fe203 NPS. The above reaction was continued for an additional 4 h in an ice bath and the final product was filtered followed by subsequent washings with 1M HCl, methanol and diethyl ether. The product obtained was dried for 48 h at room temperature. Four different weight ratios of y-FezO3: aniline monomer (1:0.1, 1:0.4, 1:06, and 1:0.8) was maintained in the synthesis procedure (from literature) with increasing concentrations of the monomer. 74 NH2 + HCI > (NH4)25203 Aniline Figure 4-1. Schematic representation of polyaniline coating of gamma-iron oxide nanoparticles. 4.1.2 EAPM Nanoparticle Characterization 4. I . 2.1 Magnetic Characterization A superconducting quantum interference device (Quantum Design MPMS SQUID) was used for magnetic characterization and room temperature hysteresis measurements of the synthesized EAPM NPs. The M-H hysteresis loop measurements for the four different EAPM samples (1:0.1, 1:0.4, 1:0.6, and 120.8) were performed at a constant temperature of 300K with the magnetic field cycling between + 20 kOe and -20kOe. The saturation magnetization (Ms) values of the EAPM NPS were determined in emu/gm and were compared with that of bare y-Fe203 NPS. The coercivity (Hg) and retentivity (MR) of the NPs were also determined from the hysteresis loop measurements in order to study the presence of super paramagnetic behavior in these samples. Zero field cooled-field cooled (ZFC-FC) measurements were performed for the 120.6 EAPM NPS for a temperature range of 5K to 300K at an applied magnetic field of 100 Oe in order to determine the blocking temperatures of the EAPM NPS. 4. 1.2.2 Electrical Characterization The electrical conductivity of the four different EAPM samples (1:0.1, 1:0.4, 120.6, and 120.8) was evaluated in the solid form. Approximately, 0.25 gm of each sample was 75 compressed into pellets of 1.5 to 2 mm thickness using a hydraulic press (Fisher Scientific, NJ) and applying a pressure of about 10,000 psi. Room temperature electrical conductivity of the compressed pellets was then measured using a Four Point Probe (Lucas/Signaton Corporation, Pro4, CA). 4.1.2.3 Structural Characterization The structural morphology of the four types of EAPM NPS were analyzed using both transmission electron microscope (TEM, Japan Electron Optics Laboratories, JEOL 100CX II) and scanning electron microscope (SEM, JEOL 6400V). The crystalline nature of the EAPM NPs was studied by selected area electron diffraction using the 200kV JEOL 2200 field emission TEM. In order to confirm the presence of the polymer, elemental analysis of the EAPM NPs was performed by X-ray energy dispersive spectroscopy using the JEOL 6400 SEM at an accelerating voltage of 20 kV and a working distance of 15 mm. 4.1.2.4 Spectral Analysis The UV-visible spectra of the four different EAPM samples were obtained using a UV-VIS-NIR Scanning spectrophotometer (UV-3101PC, Shimadzu, Kyoto, Japan). The EAPM NPs were at first dispersed in de-ionized water at a concentration of 1 mg/ml by sonication for 10 min. 1 ml of the nanoparticle suspension was transferred into a quartz cuvet (10 mm path length) and the absorbance was measured by the scanning the sample for a wavelength range of 300 to 1000 nm using a step size of 1 nm. 76 4.2 OBJECTIVE 2 Design and fabrication of an EAPM NP based immunosensor for the detection of B. anthracis spores. This objective was aimed at designing and constructing an antibody based biosensor detection system using EAPM NPS as magnetic concentrator and biosensor transducer for the rapid detection of B. anthracis (Sterne) spores as a model pathogen. 4.2.1 Biosensor Design and Data Collection The biosensor was comprised of three distinct membrane pads: sample application pad, capture pad, and absorption pad. The dimensions of the overall as well as different biosensor zones are presented in Table 4-1. The membrane pads were arranged on a polystyrene adhesive backing with silver electrodes fabricated 0.5 mm apart on the capture membrane. The application and the absorption pads were made of cellulose membranes with a flow rate 180 ml/min and the capture pad was made of nitrocellulose membrane with a flow rate of flow rate 135 sec/4cm according to the manufacturer’s specification (Millipore, Bedford, MA). For data collection, the fabricated biosensor unit was attached to an etched copper printed circuit board (PCB) and connected to a portable multimeter [BK Precision, Model 390A, range: 4009 - 40MQ, resolution: 0.1!) - 10k (2, accuracy: i (0.5 to 1.5% rdg + 4 digits), sampling rate: 2 measurements per second] with RS-232 interface to a computer. The biosensor architecture and the data collection system are shown in Figure 4-2 along with a picture of the setup. 77 .839. adenine .M no 5.398% new ESQ? cocoa—.8 «:6 EV new 239859..“ Snead—58:: woman 2.54m— ?a he eta—5.3m .N-.. 95w:— moa coaaoo 558232 .93an0 .38. . wcwhbmbom can. 53.82 meobumm Dan. SBQMU Ea 838:8... 3V 78 Table 4-1. Dimensions of the EAPM based immunosensor. Biosensor Zone Dimension (mm) Application 20 X 5 Capture 23 X 5 Absorption 17 X 5 Overall 60 X 5 4.2.2 EAPM Based Immunosensor Fabrication 4. 2. 2.1 Chemicals and Reagents Tris-HCl buffer, glutaraldehyde, polysorbate 20 (Tween 20), peptone water, and sodium phosphate (dibasic and monobasic) were purchased from Sigma-Aldrich (St. Louis, MO). Mouse monoclonal anti-anthrax IgG (clone 2C3) and polyclonal goat anti- anthrax combo IgG molecules were obtained from Chemicon International (Millipore, CA) whereas anti—goat IgG-FITC (whole molecule) was obtained from Sigma- Aldrich (St. Louis, MO). Nitrocellulose and cellulose acetate membrane pads for the biosensor were purchased from Millipore (Bedford, MA). Conductive micro tip pen used for dispensing silver electrodes were procured from Chemtronics (Kennesaw, GA). De- ionized water from Millipore Direct-Q system was used for preparing all reagents. 4. 2. 2.2 EAPM Antibody Modification The synthesized EAPM NPs were conjugated with mouse monoclonal anti-B. anthracis IgG molecules by direct physical adsorption of the antibodies onto the NPs (Figure 4-3). The EAPM NPS (100 mg/ml) were dispersed in 100 mM phosphate buffer (pH-7.4) by sonication for 10 to 20 min and mixed with anti- B. anthracis IgG (150ug/ml) antibodies. The reaction mixture was incubated for 1 h at 25°C in a rotational hybridization oven (Amerex Instruments Inc., CA). After incubation, the IgG labeled 79 NPs (immuno-EAPMS) were magnetically separated to remove the supernatant using a magnetic separator (Spherotech, IL). The resulting immuno-EAPMs were washed three times with a blocking buffer consisting of 100 mM Tris-HCl buffer (pH-7.6) and 0.1% (w/v) casein. Finally, the immuno-EAPMS were suspended in 100 mM phosphate buffer (ph-7.4) and stored at 4°C. Y + YYY Adsorption» ® Magnetic EAPM NPs Monoclonal Antibodies Antibody Labeled EAPM NPs Figure 4-3. Modification of EAPM NPs with monoclonal antibodies. 4. 2. 2.3 Biosensor Capture Pad F unctionalization Figure 4-4 illustrates the functionalization of the biosensor capture pad for polyclonal antibody attachment. The nitrocellulose biosensor capture membrane pads were first cleaned with de-ionized water followed by 10% methanol (v/v) for 45 min and dried at room temperature. The cleaned membranes were treated with 0.5 % glutaraldehyde solution (v/v) for 1 h and dried at room temperature. After drying, the glutaraldehyde activated capture membranes were coated with polyclonal goat anti-B. anthracis IgG (SOOug/ml) using a reagent dispensing module (Matrix 1600, Kinematic Automation Inc., CA) and incubated for 1 h at 37°C. Finally, the membrane surface was washed with blocking buffer consisting of 100 mM Tris-HCl (pH-7.6) and 0.1% (v/v) Tween 20 and incubated for 45 min at 37°C to inactivate the residual aldehyde groups. 80 The antibody modified membranes were dried at room temperature and stored at 4°C before use. O=CH(CH2)3CH=O Lia—n-u—HEEJ + OBCHNH2)3CH"O'—___> ' Capture Pad Glutar aldehyde Functionalized Capture Pad Crosslinker Blocking <—— <——— YNH2¥NH2 Antibody Attachment Polyclonal Antibodies Figure 4-4. Functionalization of biosensor capture pad for attachment of polyclonal _ antibodies. 4. 2.2.4 Application and Absorption Pad Preparation The application and the absorption cellulose membrane pads of the biosensor were washed with de-ionized water to remove dirt and surface residues. The membranes were then air-dried and stored in a clean container before being assembled into the biosensor. 4.2.2.5 Sensor Assembly The application, capture and the absorption pads after preparation were assembled according to the arrangement as shown in Figure 4-2 (A) with the membrane pads overlapping one another. After assembly, the biosensors were cut into 5 mm strips using a programmable shear module (Matrix 2360, Kinematics Automation Inc. CA, USA). Silver electrodes were fabricated on the capture membrane surface at 0.5 mm distances apart using silver conductive ink (Chemtronics, GA) and the biosensor strips were connected to the data acquisition systems as described as shown in Figure 4-2 (B). 81 4.2.2.6 Antibody Attachment Confirmation Studies The antibody modification of the EAPM NPs was confirmed and investigated by measuring the absorption of IgG molecules using a UV-VIS-NIR Scanning Spectrophotometer (UV-3101PC, Shimadzu, Kyoto, Japan). The UV spectrum of pure anti- B. anthracis IgG (150ug/ml) molecules and that of the unreacted IgG molecules present in the supernatant after magnetic separation of the immuno-EAPMs was studied by scanning the samples for a wavelength range of 200 nm to 400 nm using a resolution of 1.0 nm. The attachment of polyclonal anti- B. anthracis IgG molecules on the biosensor capture pad was confirmed by a Laser Scanning Confocal Microscope (Olympus Fluoview 1000). The antibody modified biosensor capture pad was labeled with secondary anti-goat IgG-FITC molecules using a working dilution of 1:320. Fluorescence was observed by excitation with 488-nm line of argon-ion laser and detection of emission using a 505 nm long pass filter. 4.3 OBJECTIVE 3 Evaluation of sensitivity and specificity of the EAPM NP based immunosensor and determination of EAPM immunomagnetic capture efficiency. This objective was aimed at determining the sensitivity of the fabricated EAPM based immunosensor in pure suspensions of B. anthracis spores and artificially contaminated food matrices, determining the immunosensor specificity in non-target bacterial cultures, and in determining the capture efficiency of EAPM NPs in immunomagnetic concentration procedures. 82 4.3.1 Bacterial Culture and Plating Characterized strains of Bacillus anthracis (Sterne) and Bacillus cereus were obtained from the Michigan Department of Community Health (Lansing, M1) for performing sensitivity studies with the immunosensor. Generic E. coli and Salmonella enterica Serovar Enteritidis (S-64) strains from the collection of the Biosensors Laboratory were used for specificity experiments. The Bacillus strains and generic E. coli were grown in trypticase soy broth (TSB, Difco) while lactose broth (Difco) was used for growing Salmonella Enteritidis. The cultures were grown in an incubator (Fisher Scientific, IA) at 37°C for 24 h and were enumerated by spiral plating in the appropriate . plating media followed by colony counting after 24 and 48h. Both the plating and colony counting were performed using automated equipments (Microbiology International, MD, USA). 4.3.1.] B. anthracis Sporulation Protocol Sporulation of B. anthracis (Sterne) cells was promoted following a previously published protocol (Pezard et al., 1991). B. anthracis vegetative cells from a 24 h enrichment culture were streaked on Nutrient Yeast Broth (NYB) agar slopes and incubated at 30°C for seven days. Sporulation was monitored by Schaeffer-Fulton staining and light microscopy. Once the sporulation was greater than 90%, the spores were harvested by suspending them in 5 ml sterile de-ionized water and incubating at 65°C in waterbath for 30 min to kill the remaining vegetative cells. The spores were collected by centrifugation at 12000 rpm for 3 min, and re-suspended in 5ml of sterile de- ionized water. Finally, the total spore count was estimated using a disposable hemocytometer (C-Chip DHC-NOl, Fisher Scientific, Pittsburg, PA) and the viable 83 spores were enumerated by microbial plating in trypticase soy agar (TSA II) with 5% sheep blood plates (BD Biosciences, MD) and counting the colonies after 16 and 36h. The spore suspension was stored at 4°C for fixture use. 4.3.2 lmmunomaggetic Concentration using EAPM NPs Following the antibody modification procedure, the antibody coated EAPM NPs (immuno-EAPMS) were added to different concentrations of B. anthracis spore suspensions (10l to 107 spores/ml) in order to have a final immuno-EAPM concentration of 20 mg/ml. The amount of EAPM NPs to be used in the immunomagnetic concentration procedure was determined from prior studies with B. cereus as target pathogen. The mixture was incubated with gentle shaking in a hybridization oven (Boekel Scientific, PA) for 10 min at room temperature. The immuno-EAPM-spore complexes formed were magnetically separated from the unbound spores, to remove the supernatant and washed twice with sterile de-ionized water before re-suspending in 10 ml sterile de-ionized water. 4.3.3 Biosensor Testing One hundred microliters of the immuno-EAPM-spore complexes obtained after immunomagnetic concentration was applied carefully to the application pad of the biosensor. The flow time of the sample from the application pad to the capture pad was approximately 1 min. Recording of the resistance signal across the biosensor electrodes was started from 2 min and carried for a total time of 6 min as determined from prior studies (Pal et al., 2007). Resistance measurements were also made before applying the sample to the biosensor to ensure that the biosensor strips were not conductive before sample application. Three trials were performed for each set of experiments. All 84 biosensor testing were performed inside a Class II Biosafety Cabinet (Model 1284, ThermoElectron Corporation, OH). 4.3.4 Biosensor Sensitivity Stud! Pure spore suspension of B. anthracis was used to assess the sensitivity of the EAPM biosensor. The B. anthracis spores stock solution was serially diluted to spore concentrations ranging from 100 to 107 spores/ml. The performance of the four different EAPM samples (1:0.1, 120.4, 110.6 and 1:08) was evaluated in the detection of B. anthracis spores prior to choosing the best nanoparticle type for the sensitivity analysis. Three different spore concentrations (101, 104 anle7 per ml) were used in this comparison study. The biosensor sensitivity was then determined using the EAPM NPs that showed the best biosensor performance in the above study in spore concentrations ranging from 100 to 107 per ml. All biosensors were baselined against a negative control solution containing EAPM nanoparticles in sterile water (concentration-20 mg/ml). 4.3.5 Biosensor Specificity Analysis Pure cultures of generic E. coli and S. Enteritidis were used for determining the specificity of the EAPM biosensor. Twenty four hour enrichment cultures of generic E. coli and S. Enteritidis were serially diluted to cell concentrations ranging from 101 to 105 CFU/ml using sterile 0.1 % (w/v) peptone water. The immuno-EAPMS were added to appropriate cell concentrations of the desired bacteria to obtain a final immuno-EAPM concentration of 20mg/ml. The immunomagnetic concentration and biosensor detection procedure was similar to that as described in Sections 4.3.2 and 4.3.3. The resistance signal generated was compared statistically with positive and negative control samples of B. anthracis spores to evaluate the biosensor specificity. 85 4.3.6 Biosensor Testing in Food Matrices The efficiency of the EAPM biosensor was tested in three different food matrices. Romaine lettuce, lean ground beef, and ultra-pasteurized whole milk, was purchased from a local grocery store. For the lettuce and beef samples, 25-g samples were weighed, mixed with 225 m1 of 0.1% (w/v) peptone water in a Whirl-Pak plastic bag, and stomached in a stomacher (Microbiology International, MD) for 1 minute. The whole milk sample was used as purchasedl Nine milliliters of the three kinds of liquid samples were thoroughly mixed with 1 ml of appropriate concentrations of B. anthracis spore stock solution. Finally, a series of 10 ml samples inoculated with B. anthracis spores at concentrations ranging from 101 to 107 spores/ml were obtained. Following the food sample preparation, immuno-EAPMs (20 mg/ml) were mixed with different concentrations of the B. anthracis spore inoculated food matrices. Immunomagnetic concentration of B. anthracis spores from the food matrices and biosensor testing was performed according to the procedure described in Section 4.3.2 and 4.3.3. EAPM NPs suspended in the liquid samples (20 mg/ml) obtained from the three different food matrices without inoculated B. anthracis spores was used as a negative control. 4.3.7 Statistical Analysis Statistical analysis of the biosensor data was performed using the SAS software (SAS, Cary, NC). The means and standard deviations of the biosensor resistance signals for. all samples were calculated on the basis of data from three replicates and the differences between the means were compared using analysis of variance (ANOVA) to a significance of 95% ((1 =0.05). Factorial ANOVA analysis (Two-way ANOVA with 86 slicing of interactions) was performed to determine the interactive effects of EAPM NP type and spore concentration as well as the double interaction of both factors on the biosensor resistance responses in the EAPM comparison study. For sensitivity study, the two-way interaction of spore concentration and time was analyzed. For specificity analysis, interactive effects of different pathogen types, cell concentrations as well as combined effect of both factors were studied using two-way ANOVA. Single factor analysis of variance was performed on the food sample data. 4.3.8 Confirmation and EAPM Capture Efficiency Calculations The immunomagnetic capture of B. anthracis spores using EAPM nanoparticles from pure spore suspension and different food matrices was confirmed by microbial plating in TSA II blood agar plates. One hundred microliters of the immuno-EAPM-spore complexes obtained after immunomagnetic concentration was used for plating. Triplets of each sample were plated. The viable spore count was enumerated after 16 to 36h. The capture efficiency (CE) of the EAPM nanoparticles was calculated using the equation: Cb CE(%)= Zn 100 where, C, is total concentration of viable spores in the sample (CPU/ml), and Cb is the concentration of viable spores bound to the EAPM nanoparticles (CFU/ml). 87 4.4 OBJECTIVE 4 Fabrication of an EAPM based electrochemical DNA biosensor for detection of B. anthracis. This objective was aimed at developing an electrochemical DNA biosensor system using EAPM NPs for target DNA concentration as well as biosensor transduction for the detection of Bacillus anthracis protective antigen (pagA) gene. 4.4.1 Biosensor Desigg and Data Collection A screen printed three electrode sensor (Gwent Group, UK) was used in the electrochemical detection of DNA targets. The sensor chip had an overall dimension of 22 X 12 mm with a circular working electrode of 4 mm diameter and a partially circular (270°) common reference and counter electrode of 1.5 mm width. Screen printed carbon acted as the working electrode while screen printed silver/silver chloride (Ag/AgCl) acted as the common reference and counter electrode on a polyester backing. The screen printed carbon and silver ink had a resistance of 50 Ohms at 12 microns, and 320 mOhms at 25 microns, respectively, according to the manufacturer’s specifications. Figure 4-5 shows the architecture of the screen printed carbon electrode (SPCE) biosensor. For data collection, the SPCE biosensor units were connected to a Potentiostat/ Galvanostat (Model 263A, Princeton Applied Research, OakRidge, TN). 88 —> Polyester Backing Reference & Counter Electrode (Common) ‘———> Working Electrode Figure 4-5. Screen printed carbon electrode biosensor. 4.4.2 Chemicals and Reagents Sodium chloride (NaCl), sodium phosphate (monobasic and dibasic), sodium acetate, phosphate buffered saline tablets (0.01M), formamide, ethylene diamine tetra acetic acid (EDTA), isopropanol, ethanol, sodium dodecyl sulfate (SDS), proteinase K, hydrochloric acid (HCl), Trizma base, ethidium bromide, gel loading solution and Streptavidin from Streptomyces avidim'i were purchased from Sigma Aldrich (St Loius, MS). AccuPrimeTM Taq DNA Polymerase system, UltrapureTM DNAse RNAse free water and UltrapureTM agarose were purchased from Invitrogen Corporation (Carlsbad, CA). QIAquick PCR purification kit was procured from Qiagen Inc. Valencia, CA. EDC (1-ethyl-3-[3- dimethylaminopropyl] carbodiimide hydrochloride) was obtained from Pierce (Rockford, IL). 89 4.4.3 Selection and Analysi_s of DNA Primers and Probes The primers pairs for PCR, and the capture and detector probes for the biosensor, were designed and selected from the protective antigen (pag A) gene, while the non- complementary sequence was selected from the capsular (cap A) gene of Bacillus anthracis (Song et al., 2005). Detailed information on the sequences of the probes and the primer pairs has been shown in Table 4-2. The specificity of the primer and the probe sequences were analyzed using the Basic Local Alignment Search Tool (BLAST). 4.4.4 Extraction, Amplification and Characterization of B. anthracis DNA 4. 4. 4.] DNA Extraction Genomic DNA was extracted from a 24 h enrichment culture of B. anthracis using a modified protocol for mammalian DNA extraction (Laird et al., 1991). One ml of the overnight culture was transferred into a 1.5 ml microcentrifuge tube and centrifuged for 10 min at 13000 rpm to precipitate the DNA. The supernatant was discarded and the remaining pellet was re-suspended in 500 pl of lysis buffer (IOOmM Tris HCl, 5mM EDTA, 0.2% SDS, 200 mM NaCl) and Sul of proteinase K solution. The tube was placed in a waterbath for 75 min at 55°C and then introduced in ice. Five hundred microliters of isopropanol was added to the tube and mixed, followed by freezing the tube for 30 min at - 80°C. The precipitated DNA was recovered by centrifugation for 15 min at 13000 rpm followed by addition of 500 pl of pre-cooled ethanol solution and centrifugation at 13000 rpm for 15 min. Finally, the supernatant was discarded and the pellet was dried to and resuspended in 100 u] of DNAse and RNAse free water. 90 ceaseeaum .. coca—bofimonm-.m @2me 0058235 no.“ 53302;»: .maoEmoq .m was .m ..m ..m 3 23a— Eimé 53> @25on 8o? 82833 02 Ea 3m-O~E .OMEAE dd 2:. AS 65280an .05 $233503. <75 @8832: 82m @8283 083 82838 =< - .mooooofiozotoozfiog 66 eoeoeeom 82 ER om 3&2 om 55 O 3 'U C o 0 8 '5 1 7 U 2 11.1 0 _# . 1:0 1:0.1 120.4 EAPM Composition 1:06 1:08 Figure 5-3. Room temperature electrical conductivity of the EAPM NPs. Table 5-2. Conductivity values measured for the EAPM NPs. EAP M C0mP9Sffi°n Fe203 Content Pellet Thickness Conductivity at ('y-Fe203:Anlllne) (Wt. %) (Hm) 300K (s /cm) 1:0 100 1488 0.00005 1:01 90 1683 0.092 104 71 1476 0.768 120.6 62 1983 1.129 1:08 55 1938 2.436 104 1110f] NPs Transmission electron microscope (TEM) images were used to study the structural morphology and size distribution of the unmodified y-Fe203 and the synthesized EAPM NPs. The TEM image of the bare y-Fe203 NPs in Figure 5-4 shows that the NPs have spherical shape and are less than 50 nm in size with an approximate size distribution of 10 to 40 nm which is consistent with the manufacturer’s specifications. Selected area electron diffraction images (Figure 5-4, inset) indicate that the 'y-Fe203 NPs are crystalline in nature. The TEM image of the 1:01 EAPM NPs (Figure 5-5, left) shows that the particles are spherical in shape similar to that of the y-Fe203 NPs. The average size distribution of the 1:01 EAPMS are smaller than 50 nm thus indicating that the NPs have very minute quantities of polymer coating on them. Crystalline nature of the 1:01 EAPMs are confirmed through the electron diffraction rings (Figure 5-5, inset). Figure 5-4. TEM and electron diffraction image of y-Fe203 NPs. 105 575$ v c _ 93: ”2.3. :3 3e: "mmz 2.31% we Bug: 353.5% E5929 ES Em: .m-m 95w:— 106 107 Figure 5-6. TEM and electron diffraction images of EAPM NPs: (left) 1:0.6 EAPM; (right) 1:0.8 EAPM. The 1:0.4 EAPM (Figure 5-5, right) aggregates show spherical morphology with an average size distribution of 100 to 200 nm and are crystalline (Figure 5-5, right inset). The 1:0.6 EAPMs (Figure 5-6, left) have an approximate size distribution of 50 to 100 nm and exhibit a chain like morphology similar to observations made by Sharma et al. (Sharma et al., 2005). The electron diffraction image (Figure 5-6, left inset) of the 1:06 EAPMs also reveal the crystalline nature of the samples. The 1:0.8 EAPMs (Figure 5-6, right) show the presence of excess polymer as compared to the 1:04 and 1:06 EAPMs and the aggregates are greater than 200 nm in size. The electron diffraction patterns of the 1:08 EAPMs (Figure 5-6, right inset) also indicate that they are mostly amorphous in nature. The TEM observations of the synthesized EAPM NPs are quite consistent with the measured magnetic and electrical parameters of the NPs. The 1:0.6 EAPM was consequently chosen over the 1:01, 1:04 and 1:08 EAPMs for performing structural characterization studies as the NPs show good saturation magnetization as well as high electrical conductivity values thus suggesting a uniform distribution of the polymer and the magnetic core in the sample. The presence of both polyaniline and y-Fe203 in the 120.6 EAPM NPs as well as its conductive nature was therefore confirmed by energy-dispersive X-ray microanalysis (EDS) of the EAPM samples using a scanning electron microscope and through UV—VIS spectral analysis. 108 .555 5 $95 b38232: o>_m._oam__rzw..o:o ”3...... £2 Shana c.9— 05 he flea—«:5 73:050.”.— .b-m charm o2 Eek m2: «mom M on 8N mam M B 2.3” 3% M 0 mg 93 M z 2.3. mix M v exp 2:82.. 3% Bus...» Eases @ m w m m .F e ”:6 on. .0 on. r o O 109 Figure 5-7 shows the EDS elemental analysis data of the 1:0.6 EAPMs. All the elements barring hydrogen present in polyaniline i.e. carbon (C), nitrogen (N), and chlorine (Cl) could be detected in the 1:06 EAPM sample. The absence of hydrogen in the EDS data was due to instrumental limitations of the EDS detector in the SEM microscope which could only detect elements having atomic numbers higher than 5. The dopant Cl atoms in the EDS data also suggest that polyaniline is present in its conductive form in the EAPM NPs. High percentages of Iron (Fe) and oxygen (O) atoms are also detected in the 1:06 EAPM NPs thus confirming presence of y-FezO3. Figure 5-8 shows the result of UV-vis spectral analysis of pure polyaniline and the 120.6 EAPM NPs dispersed in water. As observed in the figure, the characteristics peaks of pure polyaniline appear at 356 nm, 433 nm and 862 nm. The peak at 356 nm can be attributed to 7t-7t* transition of the benzenoid ring, while the peaks at 433 nm and 862 nm * are associated with polaron- 7t and a-polaron band transitions of polyaniline, respectively (Stafstrom et al., 1987;Lv et al., 2005;Dallas et al., 2006). For the EAPM NPs, characteristic absorption peaks are observed at 441 nm and 864 nm that can be related to the polaron- 7d and n—polaron band transitions of polyaniline. For the EAPM NPs, the polaron- 7t. peak shows an 8 nm shift and the n-polaron peak shows a 2 nm shift from that of pure polyaniline. These red shifts observed in the spectrum can be explained by interactions between the Fe203 nanoparticles and the polymer backbone. However, the 7:- 1t“ transition peak of the polymer is not properly distinguished in the absorbance spectrum of the EAPM NPs. The absorbance data fiirther confirms that the polymer is present in the doped (conductive) state in the EAPM NPs (Ohira et al. , 1987). 110 Absorbance 300 400 500 600 700 800 900 Wavelength (nm) [— Polyaniline —— 120.6 EAPMI Figure 5-8. UV-VIS absorption spectra of pure polyaniline and EAPM NPs. 111 5.2 @JECTIVE ; Design and fabrication of an EAPM NP based immunosensor for detection of B. anthracis spores. 5.2.1 EAPM Based Immunosensor Fabrication 5. 2. 1.1 Confirming EAPM Antibody Modification The labeling of the EAPM NPs with antibodies was confirmed by spectrophotometric studies. Figure 5-9 shows the UV spectrum of pure anti-B. anthracis IgG (lSOug/ml) molecules and that of the unreacted IgG molecules in the supernatant after magnetic separation of the immuno-EAPMS measured by a UV-VIS-NIR scanning spectrophotometer. As evident in the figure, pure IgG solution shows a characteristic peak of protein molecules at 280 nm. However, the supernatant from the immuno- EAPMs shows no characteristic peak at 280 nm, thus implying that the anti-B. anthracis IgG molecules are adsorbed onto the EAPM NPs and no unreacted IgG molecules remain in the supernatant. Physical adsorption was used for bio-modification of the EAPM NPs with IgG molecules because the conjugation procedure is simple and allows IgG molecules to retain their activity and conformation. Physical adsorption is affected by several factors, such as van der Waals interaction, electrostatic interaction, hydrophobic effect and hydrogen bonding (Zhou et al., 2004). Literature suggests that IgG adsorption occurs preferentially through Fc fragment of the molecule (Buijs et al., 1996). Hence, it can be suggested that electrostatic interactions between negatively charged Fc fragments of IgG molecules and positively charged polymer surfaces of EAPM NPs play a significant role in the physical adsorption process. 112 Absorbance l l 250 300 350 400 Wavelength (nm) —Anti- B. anthracis IgG (150uglml) — Unreacted IgG (supernatant) Figure 5-9. UV spectrum of pure anti-B. anthracis IgG molecules and unreacted IgG molecules after magnetic separation from EAPM NPs. 5.2.1.2 Confirming Biosensor Capture Pad F unctionalization The successful immobilization of polyclonal goat anti-B. anthracis IgG (500ug/m1) on the biosensor capture pad was confirmed by fluorescent images obtained from the Laser Scanning Confocal Microscope. The anti-B. anthracis IgG immobilized biosensor capture pads were modified with secondary antibodies (FITC-anti-goat IgG) using a working dilution of 1:320. Figure 5-10 (left) shows the fluorescent image of the capture pad labeled with the secondary FITC anti-goat IgG and Figure 5-10 (right) shows the image of the capture pad without the FITC label (control), observed by excitation with 113 488-nm line of argon ion laser, and the detection of emission using a 505 nm long pass filter. The bright green fluorescence emission in Figure 5-10 (left) confirms the immobilization of anti-B. anthracis IgG on the capture pad which is detected by the F ITC label of the secondary anti-goat IgG molecules. In contrast, the control image (Figure 5- 10, right) shows no significant fluorescence emission because of the absence of the secondary FITC labeled antibody in the control. The appearance of a few fluorescent dots in the control image can be explained by the tendency of the nitrocellulose membrane to fluoresce when exposed to the argon ion laser. 114 A81 cm H can 233 Czar—v Ena— Uw— anemia.“ Dir—h 3253 can 9:593 .23 £0: Ena— Ow— anemia...“ 0...: :23 can 9:593 hem—533 vogue: Macaw—E no nouns: 2.38.31: .3855 wfinnaum .525 .31.». 25E 115 5.2.2 EAPM Based Immunosensor Detection Concept Figure 5-11 is a schematic representation of the immunomagnetic separation technique and the detection procedure of the EAPM based immunosensor. The detection principle involves a sandwich immunoassay, with a capture antibody (polyclonal anti- B. anthracis IgG) immobilized on the biosensor capture pad and a detector antibody (monoclonal anti- B. anthracis IgG) conjugated with the synthesized EAPM NPs. The detector antibody conjugated EAPM NPs are first added to the target antigen contaminated samples (FigureS-l 1, step 1) to capture the antigen by applying a magnetic field (step 2). The immunomagnetically captured target antigens are then washed to remove unbound antigens and other materials (step 3) and applied to the sample application pad of the biosensor (step 4). The antigen-antibody-EAPM complex flows to the capture pad, where the antigen is anchored by the capture antibodies present on the sensor capture pad and a sandwich structure is formed (step 5). The electrically active EAPM NPs bound to the antigens in the sandwich aid in direct electron transfer across the silver electrodes thus acting as a voltage controlled “ON” switch and generating an electrical signal which is recorded (step 6) (Muhammad-Tahir and Alocilja, 2003b;Pal et al., 2007). In the absence of antigens, the sandwich complex is not generated and as a result of which the electron transfer by the EAPM NPs between the electrodes is not facilitated. 116 .353 MSES§ .M we E5982. .8.“ Lem—33:58:: comma Emdfi 2: we nits—38.50.. 23835 A Tm 0.53,.— co=Smcmm Ecmfi 928 :35 Kaboom. :ozmnocmm .chw Eta #586 _8_.aom_w :. l_____T I + I + 1 E0, 5:5 .20.. 5:5 __ : Eozfiaam 29:8 928v Eo=8__&m 238 5:3 EmEmsmmmE cozommmmeo 93 9380 .8585 cozowmmweo 28 2398 59806 acgw l m n<.o..3qmo \ mvobomxm v macaw unambcoocoo 3a 0.05). These results suggest that the effect of different EAPM types is more pronounced at 104 spores per ml. Figure 5-14 also shows that the 1:0.1 EAPM NPs can only detect the spores at the highest concentration level (107 per ml) and hence was not considered further for subsequent experiments. Similarly, the 1:08 EAPM NPs have no statistically significant effect on the biosensor resistance signals (P = 0.8227). This can be due to the greater agglomeration and larger size observed in the 110.8 NPs (in Figure5-6) which have an impeding effect on the flow of the EAPM NPs as well as the EAPM-antibody-spore complexes through the capture pad thus resulting in a high background signal which makes the control signal indistinguishable from that of the different spore concentrations. 123 The 120.6 EAPM NPs were consequently chosen for biosensor sensitivity and specificity studies. The high electrical conductivity and competent magnetic properties of the NPs confirmed by DC SQUID and four point probe measurements as well as the low variability in the resistance responses of the biosensor (Figure 5-14) at high antigen concentrations with these NPs suggests that the 1:06 EAPMS offers the best trade off between desired NP characteristics and biosensor signal generation. 500 400 — E g 300 — y a 200 _ % . .5 / a / / 100 — % / / o _ As %s A Control 10"1 10M 10"? No. of spores per ml {—l120.1l1:0.4 I1:0.6 mos. Figure 5-14. Comparison of biosensor resistance response of the four EAPM NPs at three different spore concentrations (Mean resistance :I: SD, n = 3). 124 5.3.2 Immunosensor Sensitivity Study The sensitivity study for the EAPM based B. anthracis immunosensor was performed for spore concentrations ranging from 100 to 107 spores per ml. Figure 5-15 shows the average biosensor resistance measured from three trials using the 1:06 EAPM NPs. Multi factor analysis of variance (Two-way ANOVA with time as repeated measures) to a significance of 95% (P < 0.05) was used to study the effect of cell concentrations and time on the biosensor resistance responses and to compare the differences in the resistance readings between the control and the different spore concentrations (results in Appendix A). The type 3 Tests of Fixed Effects table (Table A- 6) indicates that only spore concentration and not time has a significant effect on the biosensor resistance signal (P- value for concentration: < 0.000], P- value for time: 0.9983). The average resistance recorded for the different spore concentrations show significant statistical differences from the control solutions (P < 0.0001). The clarity in signal can be attributed to the magnetic concentration procedures in this assay which provide advantages such as low interference from other biological materials, concentration of target pathogens into smaller volumes, and reduced background signal (Kriz et al., 1998). In figure 5-15, a gradual increase in the mean resistance signal is observed as the spore concentration decreases from 107 to 100 spores/ml with an exception to the signal at 103 spores/ml. This can be explained by the fact that a higher concentration of pathogen will have a higher concentration of EAPM NPs bound to them resulting in an increased electron transfer between the electrodes and a concomitant decrease in the resistance signal. However, the resistance values recorded for the different spore 125 concentrations are not statistically different from one another. This disparities observed in the biosensor signal can be explained by a number of factors such as imperfections in biosensor fabrication, non-uniformity in EAPM samples, probabilistic interactions between the antigens and antibodies, antibody orientations on the capture pad surface and stability of the sandwich complex formed on the capture pad. The lowest B. anthracis spore concentration which produces a resistance signal statistically (P < 0.05) lower than the control is considered to be the analytical sensitivity or the detection limit of the biosensor. Since the biosensor signal at the threshold concentration of 4.2 x 102 spores per ml (estimated from spore count) is statistically significant from the control, the sensitivity of the biosensor was determined to be 4.2 x 102 spores/ml. This is an excellent sensitivity in terms of rapid identification since the ID 50 (median infectious dose) for B. anthracis has been reported to be in between 8000 and 10,000 spores and most biosensors developed for the detection of B. anthracis spores have sensitivities greater than 103 spores/ml (Cieslak and Eitzen, 1999;Krebs et al., 2006;Liu et al., 2007). At this stage of research, the biosensor can only be considered as qualitative device as the resistance signal at different spore concentrations are not statistically differentiable. Futhermore, the higher data variability observed in the control as well as at very low spore concentrations (viz. 101 and 102 spores/ml) could not be explained at this stage and would require further investigations. However, the biosensor signal demonstrates the success of the immunomagnetic capture process by the EAPM NPs, which is further confirmed by microbial plating results and capture efficiency calculations discussed in Section 5.3.5 of this chapter. 126 .890 v a: omcommou 858282 accept—U bag—mama 88:2: 23 2: 226 2228595 “Shaman Am u : .Qm n." 3:323.— naoav 2.3.323 .N no 322.253 93% 9::— 3 «2.2.3.. 99:328.. homaomoaafifim .323 534% .me 9:.»5 .5. con 3.3% “.0 .oz 95.. m 8.5mm 0...“; 8..on o...a.> H 2:80 0..on . . 93a...» .39.. .93. .25ch 2...). 0.2.3 358.. 052:3. .33. 258 85... 89... 6038.20 3.8% 26335» .M 5.3 6035:3239 .92. 259% .25 3...: 9.2.3 .333. 2..an.— :. aziménsfifi. no any. 55.950 9:595 dim 0.2.5... 137 5.4 OBJECTIVE 4 Fabrication of an EAPM based electrochemical DNA biosensor for detection of B. anthracis. 5.4.1 Selection and Analysis of DNA Primers and Probes B. anthracis Sterne strain, a strain with attenuated virulence, was used as the model organism in this study. Virulent strains of B. anthracis have two virulent plasmids: pXOl (184 kb), which encodes toxins and pX02 (95 kb), which encodes the capsule (Reif et al., 1994). The B. anthracis Sterne strain lacks the pX02 plasmid encoding the poly-y-D glutamic capsule which is a major determinant of virulence in anthrax. For this reason, the PCR primers and the DNA probes were selected from a fragment of the pag A gene, which is present in the plasmid pXOl. The specificity of the selected probes and primers were checked using the BLAST analysis tool. BLAST results indicated that both the probe and primer pairs had a high degree of specificity (AB value: 8e-19, Max ident: 100% for the detector probe; AE value: 3e-7, Max ident: 100% for the capture probe; AE value: 0.43, Max ident: 100% for the primer pairs) for the B. anthracis Sterne strain plasmid pXOl pag A gene. The only Bacillus species with which the sequences showed high similarity was B. cereus (B. cereus strain 69241 plasmid pBCXOl, Max ident: 100%). 5.4.2 Extraction, Amplification and Characterization of B. anthracis DNA Genomic DNA was extracted from overnight cultures of B. anthracis grown in TSB broth using a procedure for mammalian DNA extraction. The concentration of the double stranded DNA (dsDNA) after the extraction process was determined to be 326.6 ug/ml by measuring the absorbance at 260 nm in a spectrophotometer following which the 138 DNA was diluted according to the requirements of PCR reaction. The PCR amplification reaction was optimized with respect to the concentration of the reactants, the melting temperature (Tm) and the amplification time in order to obtain the best amplification of targets in the shortest possible time. Table 5-5 shows the optimized concentrations of the reactants used in the PCR reaction. Table 5-5. Optimum concentration of PCR components. :$;'§.:::::;':;:::.: Buffer 1X dNTP mix 0.2mM MgClZ 1.5mM Accuprime Taq DNA Polymerase 1 ul Forward Primer 100 nM Reverse Primer 100 nM DNA Template 100 ng The PCR amplification reaction was optimized and run under the following conditions: 95°C for 2 min; 35 cycles of 95°C for 30 s, 55°C for 303, 72°C for 605; and 72°C for 5 min in the DNA thermal cycler. The PCR products obtained were first purified using the QIAquick kit and then characterized by gel electrophoresis and spectrophotometric measurements. 139 BA1 Figure 5-18. Gel Electrophoresis of PCR amplified Bacillus anthracis. (M- Marker, BA1 and BA2 — B. anthracis, NC— Negative Control) Figure 5-18 shows the result of the PCR reaction after purification on 2.5 % agarose gel. The appearance of distinct bands in lanes BA1 and BA2 just above the 100 bp band of the marker confirms the length of the PCR product (120 bp). The appearance of bands in lanes BA] and BA2 and no bands in lane NC also confirms the successful optimization of the PCR reaction conditions along with the PCR purification process. The dsDNA concentration in the purified PCR products (BA1 and BA2 pooled together), when tested in the NanoDrop Spectrophotometer, ranged in between of 7.0 and 12.0 ng/ul with Azm/zgo ratios in between 1.63 and 2.38, thus indicating the high quality of the PCR product. 140 5.4.3 Confirming EAPM -DNA Probe Modification The biomodification of the EAPM NPs with DNA probes (Ph-PRO) was confirmed by fluorescence and spectrophotometric studies. Figure 5-19 shows the fluorescence intensity measurements of the pure 6—FAM labeled Ph-PRO probe solution [22.5uM] and that of the unreacted probes remaining in the supernatant after magnetic separation of the PRO-EAPMS (probe labeled EAPMs) by excitation at 495 nm and detection of emission at 520 nm for four different EAPM NP concentrations (0.1, 1, 10 and 20 mg/ml). As evident in the figure, the 6-FAM labelled Ph-PRO probes (pure probe) show the highest fluorescence signal since there are no EAPM NPs present in the solution. In comparison, the supernatant obtained from magnetic separation of the PRO-EAPMS after the labeling study using different EAPM NP concentrations show significantly lower fluorescence signal than that of the pure probe thus indicating the attachment of the probes to the NPs. A linear decrease in the fluorescence signal is also observed as the NP concentration increases from 0.1 to 20 mg/ml which is expected since an increased EAPM NP concentration would result in a greater number of attachment sites (terminal amine groups of polyaniline) for the phosphorylated probes. The single-stranded DNA (ss- DNA) concentration measurements of the pure probes and the supematants using the NanoDrop spectrophotometer further confirmed the attachment of the probes to the EAPM NPs. As observed in figure 5-19, the ss-DNA concentration for the pure probe is 385.2 ng/ul, whereas for the supematants from PRO-EAPMS, the ss-DNA concentration decreases and is in the range of 310.2 and 0 ng/ul. From the fluorescence and the absorbance results, an EAPM concentration of 1 mg/ml was chosen for subsequent hybridization studies, since at this concentration, the EAPM NPs were saturated with 141 adequate concentrations of the DNA probes required for hybridization with the target DNA. 8.00E+05 385.2 ng/ul 9? 6.00E+05 - tn 5 I. § 310.2 ng p §4waw~ ca 0 m 2 2 166.2 ng/pl E 2.00E+05 — 55.2 ng/pl 0 ng/pl 0 ng/pl 0 00E+00 “ I I 1 Pure 0.1mg/ml 1mg/ml 10 mg/ml 20 mg/ml Control Probe EAPM Concentration (mg/ml) Figure 5-19. Fluorescence signal of pure Ph-PRO probes and unreacted Ph-PRO probes after magnetic separation from the EAPM NPs. 142 5.4.4 EAPM Based Electrochemical DNA Biosensor Detection Concept Figure 5-20 is a schematic representation of the detection mechanism of the EAPM based electrochemical DNA biosensor. The detection principle involves an electrochemical sandwich assay engaging a detector DNA probe and a. capture DNA probe. The detector probe is labeled with EAPM NPs whereas the capture probe is labeled with biotin. Both the EAPM labeled detector probe and biotinylated capture probe are first added to the target sequence, where the target undergoes sandwiched hybridization with both probes. The EAPM-target-biotin DNA hybrids thus formed are separated from other noncomplementary sequences and unreacted DNA by magnetic separation of the EAPM NPs. The EAPM-target-biotin DNA hybrids are then added directly to the surface of streptavidin modified screen printed electrodes for anchoring the hybrids on the electrode surface by streptavidin-biotin interactions. After a short incubation period, the electrode surface is washed to remove the excess EAPM NPs and the unbound DNA hybrids. The target DNA is finally detected on the SPCE biosensor surface by the redox activity of the EAPM NPs using cyclic voltammetry. 143 .23» Mom 22255» .M 25980.. 3.. 32.28... <29 18.82.2500... .522. 223$ 2.. he 5.3.5839. 9.382.9m .cN-m 92%.... 320.85 womw uoEUOE SUSSmem mU_..D>r_ <20 C_u0_muumo.m._.s§n_: <20 .0 8:983 0:885. - . m :0 m mew o QZcfimkwucgh cocwyflmmm 2.02. 2338 2.9:. 8820.6 . . . . 822.6265 382 oz 2&5 I e 3 1| WW+@@ <20 69m» hoot to! “‘0! \ol 144 5.5 OBJECTIVE 5 DNA hybridization on EAPM NPs and sensitivity and specificity evaluation of the EAPM based electrochemical DNA biosensor. 5.5.1 Sandwiched Hybridization of DNA Targets on EAPM NPs 5. 5. 1. 1 Determination of Hybridization Conditions The EAPM labeled detector probes (PRO-EAPMS) were hybridized with the target DNA obtained from the PCR product. The optimum temperature required for the hybridization reaction was determined by fluorescence assays using 6—FAM labeled PCR products (Xexcitauon: 495 nm, Remission: 520 nm). To determine the optimum temperature, the hybridization reaction was performed for 1 hour at three different temperatures: at 25 °C (room temperature), at 45 °C (25 °C below the melting temperature, Tm, of the Ph- PRO probe) and at 55°C (15 °C below the Tm of the probe). Figure 5-21 shows the normalized fluorescence intensity of the DNA targets obtained from PCR reaction before and after hybridization with the PRO-EAPMs at the three different temperatures. At each temperature, the fluorescence signal of the PCR target is normalized against the PCR negative control to eliminate the effect of any background fluorescence. It is evident from figure 5-21 that the highest decrease in fluorescence signal for the PCR target DNA is observed at 45 °C with the reduction in normalized fluorescence being 11448 :l: 106.4 counts/0.1 s. This is expected since a hybridization temperature 25 °C below the Tm (Tm of Ph-PRO probe is 704°C) allows the maximum rate of hybridization (Mehlen et al., 2004). For the hybridization studies at 25 °C and 55 °C, the reduction in normalized fluorescence are only 5030 2!: 117.5 and 448 i 110.9 counts/0.1 s, 145 respectively. Therefore, the hybridization temperature was kept at 45 °C for subsequent studies. 14000 E 2' 12000 ~ 2: ‘3‘ c 10000 — 9. 3 c 8000 ~ 0 U i o 6000 - 2 LL 3 4000 — is a E 2000 ~ 0 z 0 .. 250 45C 550 Hybridization Temperature (C) I Before Hybridization C! After Hybridization] Figure 5-21. Normalized fluorescence signal of PCR DNA targets before and after hybridization at different hybridization temperatures. In order to ensure a fast hybridization reaction between the target DNA and the Ph- PRO probe labeled EAPM NPs (PRO-EAPMS), the hybridization time was optimized by fluorescence assays using 6-FAM labeled PCR products. The DNA hybridization reaction was performed for 30 min, 1 h and 2 h at 45 °C. Figure 5-22 shows the normalized fluorescence intensity (background subtracted) of the PCR targets before and after hybridization with the PRO-EAPMS at three different times. As observed in the 146 figure, the decrease in the fluorescence signals of the PCR targets after the hybridization event are almost the same for the three hybridization times trialed. The reduction in the normalized fluorescence values for hybridization times of 30 min, 1 h, and 2 h are 10862 :1: 107.6, 10380 i 64.5, and 10226 :t 300.8 counts/0.15, respectively. Thus, the hybridization time was limited to 30 min in the succeeding studies. 14000 E Q 12000 — a E 0 10000 — 9 8 c 8000 ~ 0 0 i o 6000 - 2 LI. 3 4000 ~ g E O 2000 — I z 0- — - 30 60 120 Hybridization Time (min) I I Before Hybridization CI After Hybridization J Figure 5-22. Normalized fluorescence signal of PCR DNA targets before and after hybridization at different hybridization times. 5.5.1.2 Confirming Sandwiched Hybridization on EAPM NPs The sandwiched hybridization of the PCR targets with the EAPM labeled detector probe (PRO-EAPMS) and the biotinylated capture probe (PRO-Bio) was confirmed by 147 fluorescent assays using 6-FAM labeled PRO-Bio probes (kexdtation: 495 nm, Langston: 520 nm). The hybridization temperature and time was fixed at 45°C and 30 min, respectively, to ensure maximum hybridization of the targets with the detector probes. Three different concentrations of the PRO-Bio probes (luM, SuM, and IOuM) were trialed in the sandwich hybridization event. Figure 5-23 shows a comparison of the change in fluorescence signal of the PRO-Bio probes in the presence and absence of the PCR targets during the sandwiched hybridization. 3.E+06 2.E+06 — bf 71? “I ". flit» E M : 2.E+06 ‘ ‘11:: e an a» ”in. 2 1.E+06 — 3» ‘4 g 3 z u.- 5.E+05 S 3: ”i 2? at . _J—i-—-1 O.E+00 2 ‘ I L I Pure Probe Target No Target Sample {Mum D5pM mow]; Figure 5-23. Change in fluorescence of PRO-Bio probes in the presence and absence of PCR targets during sandwiched hybridization. 148 As evident in Figure 5-23, the fluorescence signal of the PRO- Bio probes reduces by an order of magnitude in the presence of the PCR targets during the dual hybridization process for all three probe concentrations. This signal reduction confirms the successful sandwiched hybridization of the biotinylated capture probes with the PCR targets. In the absence of the PCR targets, a small decrease in fluorescence is observed at all capture probe concentrations. This can be attributed to mild bleaching of the fluorescent labels during the stringent hybridization and washing procedures. From the above results a PRO-Bio probe concentration of SuM was fixed for the successive detection experiments and the resulting EAPM-target-biotin DNA hybrids from the sandwiched hybridization were used directly for electrochemical detection. 5.5.2 Electrochemical Detection Using SPCE Biosensor 5. 5. 2.1 Electrochemical Characterization of EAPM NPs Electrochemical characterization of the synthesized EAPM NPs was performed on the screen printed carbon electrodes (SPCEs) using cyclic voltammetry (CV) before proceeding to the detection of EAPM captured DNA hybrids. Figure 5-24 shows the cyclic voltammogram of EAPM NPs in 0.1 M HCl scanned from -O.4 V to 1.0 V at a scan rate of 20 mV/s. Two stable redox peaks are observed in the voltammogram that can be related to the presence of the polymer, polyaniline, in the EAPM samples. The anodic peaks at 0.12 V and 0.59 V correspond to the switching of the leucoemeraldine base to emeraldine salt and emeraldine to pemigraniline salt respectively (Arora et al., 2007;Gospodinova et al., 1996). The corresponding cathodic peaks for the reduction process are observed at 0.53 V and -0.07 V, respectively. 149 Figure 5-25 shows the effect of scan rate on the rate of electron transfer to the SPCE electrodes by the EAPM NPs. The two characteristic redox peaks of the EAPM NPs are visible at all scan rates. With an increase in the scan rate, the anodic peak potential of the EAPM NPs is shifted to a more positive value and the cathodic peak potential is shifted to a negative direction. Figure 5-25 (inset) shows a plot of the dependence of the anodic peak current and the cathodic peak current versus the scan rate. As observed in the figure, the peak currents exhibits a linear relationship with increasing scan rate thus suggesting surface controlled behavior and diffusion controlled system (Bard and Faulkner, 2001). However, the ratio of the peak currents at different scan rates [pa/1pc 76 1 which indicates a quasi-reversible chemistry of the EAPM NPs and the SPCE electrode process (Bard and Faulkner, 2001;Ram et al., 1999). 150 at»... 2 a 6: 2:. E 8% 2.3 3 E: 8.5 .5 £2 2.2m 3.... ”sausage? 2.95 +3 25:. E om§o> 3 mo 3 no no to to- no- no- N ,.t i a z _ z mOIwo.ml mocc< . moms? t - f 8+mo.o mom? 2858 _ mowed (v) wanna 151 .83 ~38 .m> 23...:5 :3.— 95233 can 9:5...“ no 32m .32: .m\>E can "Ea c2 .2: .cm 6N he 83.. ..an «a _U= $2.: 5 £2 Em o=9m0 .mutm 0.5»:— E $5.; 2 no No no no to to- no- no- » z k z _ bl h _ VOIMO.NI . or e w E ....... .. .. . ................. . 3 8m 3%. F- . t . , e2: om» ,1 .. .. ass 2: It! .///- .. ,\ // Ema. F- 2%. .....».,.g.,..t...,_. \ t x . flit; , w\>E om . t t t . t/x.-- \ // a2: om ll - memos- 8+mo.o no to r. 3 woman . com- x/AmxxitI/ttuw tttttt I: \xwxjs . . , cor- I if?) t \t .t... :1... votmoé d .. w .. o M . .. 3mm.» I .H . .. x I V . .-. . . .. .. oo_. It .......... . vowed O O N (v) tuauno 152 5. 5. 2.2 Detection of EAPM- T arget-Biotin DNA Hybrids Following the sandwiched hybridization of the PCR DNA targets with the EAPM- PRO probes and the PRO-Bio probes, the EAPM-target—biotin DNA hybrids were directly detected electrochemically on the SPCE electrodes. Figure 5-26 shows the electrochemical response of the EAPM captured target DNA hybrids from undiluted PCR products (concentration-7.3ng/ul) on the SPCE electrode. The electrochemical responses of the bare SPCEs and streptavidin modified SPCEs were compared as control. As observed in the figure, the cyclic voltammogram of the EAPM captured targets shows the two characteristic redox peaks of the EAPM NPs that were absent in the control. The anodic peak potentials are located at 0.12 V and 0.59 V, whereas the cathodic peak potentials are located at 0.53 V and -0.09 V, similar to the CV response obtained with just EAPM NPs in 0.1M HCl. The CV response demonstrates the fact that the EAPM NPs maintained their native electrochemical behavior after the probe labeling and sandwiched target hybridization procedures. The appearance of the EAPM redox peaks in the CV response of the PCR targets also confirms the successful capture of the EAPM- target DNA-biotin hybrids and detection of the EAPM NPs on the streptavidin modified SPCE electrodes. 153 ASE cu «a .0: 92.: E Hon—m comm—5E £23392: he 3:“ £95 2.3 me £95 .8 metab— .3 F mo co to No 0 No- to- . _ . _ _ _ _ 8mm. 7 motmo. T ootmoht oo+wo.o (v) wanna memo...“ 593 «Ga .18: wk t mo-m_o.e t moflme mom—ON 154 5.5.3 SPCE Biosensor Sensitivity Study The sensitivity of the EAPM based SPCE biosensor was estimated in different concentrations of the PCR target DNA obtained by serially diluting the PCR product. Figure 5-27 shows the electrochemical response (cyclic voltammograms) of the biosensor in PCR target concentrations ranging from10 ng/ul to 1 pg/ul and the control. The control consists of just EAPM NPs with no PCR target (shown as O ng/ul in the figure). As observed in the CV response, the two characteristic redox peaks of the EAPM NPs are present in different concentrations of the PCR target. A gradual decrease in the intensity of the redox peaks is also noted with a decrease in the target concentration. This is expected since a lower concentration of the target would result in lower concentration of the EAPM-target DNA-biotin hybrids bound on the streptavidin coated electrode. The presence of the EAPM redox peaks at lower target concentrations also shows that the EAPM NPs are capable of generating redox signals at very low concentrations. Although both redox peaks of EAPM NPs are evident at lower target concentrations, the anodic peak current (oxidation peak) at an anodic peak potential of 0.59V was chosen for further analysis due to the minimal shift observed in the peak current in repeated observations at this peak potential. Figure 5-28 shows a plot of the anodic (oxidation) current in the potential range of 0.3 V to 0.7V obtained from the cyclic voltammograms of the different PCR target concentrations and the control. 155 .m\>E a he 85.. :58 a «a .0: $2.: 5 <75 333 MUA— mo 33223359 “5.5:? new 32.085 MOhm woman Em 3 mo No no no to to- no- no- a z _ _ _ z _ WOIMOO.VI . Snood- é? o . momoow- 39 god r momooe- 0 m . 8+de W m - momoo.» . momooa new» mod .39. or . momoow . lit mom—00¢ 156 -4.60E-06 - -7.0E-05 i .. 10 ng/pl C -4.10E-06 - g ‘3, i —3.60E-06 - '7; 5 g '3-1054’5 ‘ 2.0505 . r . g 0.2 0.4 0.6 0.8 E ~2.60E-06 ‘ Voltage (v) U .2 - _ ‘5 2.10E-06 109/pl C < -1.60E-06 . °‘1n9’“' - - - - 0.01ng/pl -1.1OE-06 — —-—-0.001ng/p| a.» - Ong/ul -6.00E-O7 . -1.00E-O7 . , fl 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Potential (V) Figure 5-28. Anodic (oxidation) peaks of PCR target concentrations ranging from 1 ugly] to 0 ng/ul. Inset- anodic peak of 10 ng/ul PCR target (obtained from cyclic voltammograms). As evident in Figure 5-28 (Inset), the anodic peak current at 0.59 V is maximum (62uA) for the highest PCR target concentration (10ng/ul) and decreases from 62uA to 2 uA as the PCR target concentration decreases from 10 to 0.01 ng/ul. The anodic peak currents for the control and for a target concentration of 0.001 ng/ul are 0.84 and 0.29 uA, respectively. The higher current values of the control than 0.001 ng/ul of the target indicate that there is no detection at this concentration and the peak current signal is generated by the nonspecifically adsorbed EAPM NPs on the electrode surface for both C3868. 157 1.00E-04 1.00E-05 - 1.00E-06 — Anodic Peak Current (A) 1.00E-07 — 0 0.001 0.01 0.1 1 10 DNA Concentration (ng/ul) Figure 5-29. CV mediated anodic peak current at different PCR target concentrations from three experimental trials (mean current :1: SD, n = 3). Figure 5-29 shows the mean anodic peak current of different concentrations of PCR targets obtained from three experimental trials of the EAPM based SPCE biosensor. It is evident from the above figure that the mean anodic peak current for the SPCE biosensor increases with increasing target concentration. A linear increase in the anodic current is observed for target concentrations between 0.01 and 1 ng/ul which is followed by an exponential increase in the anodic current at the highest target concentration of 10 ng//ul. It is also noted that the lowest PCR target concentration for which the mean anodic peak current can be differentiated from that of the control is 0.01 ng/ul. Based on these results 158 from the cyclic voltammetry experiments, the lowest detection limit of the EAPM based SPCE biosensor is determined to be 0.01 ng/ul of PCR DNA. To confirm the sensitivity results obtained from CV response, the anodic peak current signal of the biosensor for the different PCR concentrations were statistically analyzed by one-way AN OVA using the SAS software (results in Appendix A). From the tests of fixed effects table (Table A-21), it is evident that the PCR DNA concentrations have significant effect (P < 0.0001) on the anodic peak current signal. The table of estimates (Table A-22) further shows that the peak current values for DNA concentrations ranging from 10 ng/ul to 0.01 ng/ul are significantly different from that of control, whereas, the peak current at 0.001 ng/ul of DNA is not significantly different from the control (P = 0.6943). Thus, the statistical analysis results affirm the SPCE biosensor sensitivity to be 0.01 ng/ul of PCR DNA. Furthermore, the differences of least squares means table (Table A-24) shows that the anodic peak current signal is significantly different for different DNA concentrations (P S 0.05) thus indicating that the SPCE biosensor is quantitative for the range of DNA concentrations tested. Results of the SPCE DNA biosensor conclude that EAPM nanoparticles can be successfully used as a redox indicator in biosensor detection systems. This is a unique approach for DNA detection that has not been investigated in current literature and its potential needs to be explored in future studies. Other studies have reported DNA detection on polyaniline modified electrodes using redox indicators such as methylene blue and duanorubicin (Arora et al., 2007;Chang et al., 2007a). 159 5.5.4 Specificity Analysis The specificity of the detector probe labeled EAPM NPs (Pro-EAPMS) was evaluated by fluorescence assays using 6-FAM labeled noncomplementary sequence (NC). The NC sequence was designed from the cap A gene of Bacillus anthracis. Theoretical alignment of the NC sequence with the detector probe sequence using the BLAST analysis tool indicated no significant similarity between the sequences. Figure 5-30 shows fluorescence results before and after hybridization of the Pro- EAPMs with the 6-FAM labeled NC sequences at three different concentrations (luM, SuM and 10 uM). As evident in the figure, a small decrease in the fluorescence signal is observed for the 6-FAM labeled NC sequences at the three concentrations tested after hybridization with the Pro-EAPMS. This can be attributed to the occurrence of small amounts of cross reactions between the detector probe and the NC sequence as both sequences were designed from the same microorganism. Although theoretical results (BLAST analysis) indicated highly dissimilar sequences but such ideal conditions are hard to achieve in real samples. However, the reduction in fluorescence signal for the NC sequences were not significant as none of the concentrations showed a decrease in fluorescence intensity by an order of magnitude. Therefore, specificity studies with different noncomplementary sequences designed from different microorganisms are required in future for further specificity analysis. 160 1.0E+07 1.0E+06 T 1.0E+05 - 1.0E+04 ~ 1.0E+O3 - Fluorescence Intensity 1.0E+02 - 1.0E+01 ~ 1.0E+00 — 1pM 5uM 10pM Non-Complementary DNA Concentration I Before Hybridization C! After Hybridization Figure 5-30. Change in fluorescence intensity of non-complementary sequences before and after sandwiched hybridization on EAPM NPs. 161 CHAPTER 6: CONCLUSION AND FUTURE RESEARCH In this dissertation the potential of electrically active polyaniline coated magnetic nanoparticles (EAPM) as a novel nanostructured transducer and as a magnetic concentrator was explored for biosensor applications. The five major objectives of this research describe synthesis and characterization studies of the EAPM nanoparticles, fabrication and implementation of an EAPM nanoparticle based immunosensor, and fabrication and implementation of an EAPM based DNA biosensor. The EAPM nanoparticles were synthesized by a chemical polymerization method using y-Fe203 nanoparticles and the monomer aniline. DC Squid magnetism and Four point probe conductivity studies showed that the synthesized nanoparticles were in the ferromagnetic regime and had room temperature electrical conductivity as high as 2.4 S/cm. TEM analysis indicated that the nanoparticle size was related to polymer concentration and SEM EDS analysis as well as UV-VIS spectral analysis further confirmed the presence of the polymer in the EAPM nanoparticles in the doped or conductive state. An integrated antibody based biosensor (immunosensor) using EAPM nanoparticles as an immunomagnetic concentrator and transducer was designed and fabricated. The fabricated biosensor was implemented in the detection of B. anthracis Sterne spores. The detection process was fast and included a magnetic concentration time of 10 min and signal detection time of 6 min. A comparison study showed that the 120.6 EAPM nanoparticles performed better in the detection system and hence was chosen for successive experiments. The EAPM based immunosensor had a lower detection limit of 4.2 X 102 B. anthracis spores/ml in pure spore suspensions that was well below the 162 reported ID 50 for B. anthracis. The immunosensor sensitivity was also evaluated in food matrices and was found to be 4.2 x 102/4.2 X 103 spores/m1. Specificity analysis indicated no cross reactivity in the sensor signal with non-target pathogens such as E. coli and S. Enteritidis. Capture efficiency (CE) experiments using EAPM nanoparticles confirmed successful immunomagnetic concentration of the target pathogens with the highest CE value of 97% in concentrations as low as 101 spores/ml. A DNA biosensor was developed based on sandwiched hybridization of DNA targets on probe labeled EAPM nanoparticles and electrochemical detection of the hybridization process through the redox signal of EAPM nanoparticles. The EAPM based DNA biosensor was fabricated and implemented in the detection of B. anthracis pag A gene. Fluorescence assays confirmed the sandwiched hybridization of PCR amplified DNA targets with EAPM labeled detector probes and biotinylated capture probes. The EAPM DNA biosensor was able to detect B. anthracis DNA with a lower detection limit of 0.01 ng/ul of PCR amplified targets. The electrochemical detection process was completed in 60 min. In summary, this research shows the feasibility of using EAPM nanoparticles for magnetic concentration and biosensor transduction in more than one detection system. Table 6-1 compiles the specifications of the two different EAPM nanoparticle based biosensors that were developed. 163 Table 6-1. Specifications of EAPM based biosensors. Parameters Immunosensor DNA sensor Detection mechanism Charge transfer resistance Redox activity Target B. anthracis spores B. anthracis pag A gene Lower detection limit 4.2 X 102 spores/ml 0.01 ng/ul PCR DNA Range Of Detection 102 to 107 spores/ml 10 - 0.01 ng/ul Detection time 16 min 60 min Specificity Specific Specific Future research should be directed toward a number of different activities: (1) synthesizing uniform EAPM nanostructures with excellent electrical and magnetic properties and minimum nanoparticle aggregation; (2) advancing the EAPM based immunosensor toward quantitative evaluation of samples; (3) implementing the DNA based biosensor in detection of genomic DNA for field based applications; (4) exploring different techniques for electrochemical detection of EAPM nanostructures; (5) advancing EAPM nanostructures as an efficient magnetic concentrator; (6) developing EAPM nanostructure based biosensors for different targets , and (6) developing low-cost EAPM based magnetic detection devices. 164 APPENDIX A: STATISTICAL ANALYSIS RESULTS A.l ANOVA ANALYSIS OF DIFFERENT EAPM BASED IMMUNOSENSORS The Mixed Procedure Table A-1. Type 3 Tests of Fixed Effects Num Den Effect DF DF F Value Pr > F Concentration 3 128 3 1 .73 <.0001 Sample 3 128 22.89 <.0001 Sample*Concentration 9 128 5.47 <.0001 Table A-2. Tests of Effect Slices Num Den Effect Conc. Sample DF DF F Value Pr > F Sample*Concentration 10"1 3 128 14.29 <.0001 Sample*Concentration 10"4 3 128 8.88 <.0001 Sample*Concentration 1007 3 128 0.43 0.7351 Sample*Concentration Control 3 128 15.71 <.0001 Sample*Concentration l : 0.1 3 128 17.14 <.0001 Sample*Concentration l : 0.4 3 128 5.81 0.0009 Sample*Concentration 1 :0.6 3 128 24.89 <.0001 Sarnple*Concentration 1 :0.8 3 128 0.3 0.8227 165 Table A-3. Estimates Label Estimate Stgndard DF t Value Pr > |t| rror 10"1 vs. Control for sample 0.1 38.8 35.7967 128 1.08 0.2804 10"4 vs. Control for sample 0.1 -27.9444 35.7967 128 -0.78 0.4365 10"7 vs. Control for sample 0.1 -198.67 35.7967 128 -5.55 <.0001 10"] vs. Control for sample 0.4 -31.4889 35.7967 128 -0.88 0.3807 10"4 vs. Control for sample 0.4 -81.5222 35.7967 128 -2.28 0.0244 10"7 vs. Control for sample 0.4 -139.53 35.7967 128 -3.9 0.0002 10"1 vs. Control for sample 0.6 -105.82 35.7967 128 -2.96 0.0037 10"4 vs. Control for sample 0.6 -244.7 35.7967 128 -6.84 <.0001 10"7 vs. Control for sample 0.6 -270.57 35.7967 128 -7.56 <.0001 1001 vs. Control for sample 0.8 -21.8556 35.7967 128 -0.61 0.5426 10"4 vs. Control for sample 0.8 -26.0778 35.7967 128 -0.73 0.4676 1007 vs. Control for sample 0.8 -31.9667 35.7967 128 -0.89 0.3735 Table A-4. Least Squares Means Effect Conc. Sample Estimate Stgndard DF t Pr >|t| rror Value Sample*Conc. 10"1 0.1 308.4 25.3121 128 12.18 <.0001 Sample*Conc. 10"4 0.1 241.66 25.3121 128 9.55 <.0001 Sample*Conc. 10"7 0.1 70.9333 25.3121 128 2.8 0.0059 Sample*Conc. Control 0.1 269.6 25.3121 128 10.65 <.0001 Sample*Conc. 10"1 0.4 210.46 25.3121 128 8.31 <.0001 Sample*Conc. 10"4 0.4 160.42 25.3121 128 6.34 <.0001 Sample*Conc. 10"7 0.4 102.41 25.3121 128 4.05 <.0001 Sample*Conc. Control 0.4 241.94 25.3121 128 9.56 <.0001 Sample*Conc. 10"] 0.6 233.6 25.3121 128 9.23 <.0001 Sample*Conc. 10"4 0.6 94.7222 25.3121 128 3.74 0.0003 Sample*Conc. 1007 0.6 68.8556 25.3121 128 2.72 0.0074 Sample*Conc. Control 0.6 339.42 25.3121 128 13.41 <.0001 Sample*Conc. 10"1 0.8 78.7 25.3121 128 3.11 0.0023 Sample*Conc. 10"4 0.8 74.4778 25.3121 128 2.94 0.0039 Sample*Conc. 10"7 0.8 68.5889 25.3121 128 2.71 0.0077 Sample*Conc. Control 0.8 100.56 25.3121 128 3.97 0.0001 166 Table A-5. Differences of Least Squares Means Effect Sample Cone. __Conc. Estimate 3;de DF Value Pr > M Sample*Conc. 0.1 10"] 10"4 66.7444 35.7967 128 1.86 0.0645 Sample*Conc. 0.1 10"] 10"7 237.47 35.7967 128 6.63 <.0001 Sample*Conc. 0.1 10"] Control 38.8 35.7967 128 1.08 0.2804 Sample*Conc. 0.1 10"4 10"7 170.72 35.7967 128 4.77 <.0001 Sample*Conc. 0.1 10"4 Control —27.9444 35 .7967 128 -0.78 0.4365 Sample*Conc. 0.1 10"7 Control -198.67 35.7967 128 -5.55 <.0001 Sample*Conc. 0.4 10"] 10"4 50.0333 35.7967 128 1.4 0.1646 Sample*Conc. 0.4 10"] 10"7 108.04 35.7967 128 3.02 0.0031 Sample*Conc. 0.4 10"] Control -31 .4889 35 .7967 128 -0.88 0.3807 Sample*Conc. 0.4 10"4 10"7 58.01 11 35.7967 128 1.62 0.1076 Sample*Conc. 0.4 10"4 Control -81.5222 35 .7967 128 -2.28 0.0244 Sample*Conc. 0.4 10"7 Control -] 39.53 35.7967 128 -3.9 0.0002 Sample*Conc. 0.6 10"] 10"4 138.88 35.7967 128 3.88 0.0002 Sample*Conc. 0.6 10"] 10"7 164.74 35.7967 128 4.6 <.0001 Sample*Conc. 0.6 10"] Control -105.82 35.7967 128 -2.96 0.0037 Sample*Conc. 0.6 10"4 10"7 25.8667 35.7967 128 0.72 0.4712 Sample*Conc. 0.6 10"4 Control -244.7 35.7967 128 -6.84 <.0001 Sample*Conc. 0.6 10"7 Control -270.57 35.7967 128 -7.56 <.0001 Sample*Conc. 0.8 10"] 10"4 4.2222 35.7967 128 0.12 0.9063 Sample*Conc. 0.8 10"] 10"7 10.1111 35.7967 128 0.28 0.778 Sample*Conc. 0.8 10"] Control -2] .8556 35.7967 128 -0.6] 0.5426 Sample*Conc. 0.8 10"4 10"7 5.8889 35.7967 128 0.16 0.8696 Sample*Conc. 0.8 10"4 Control -26.0778 35.7967 128 -0.73 0.4676 Sample*Conc. 0.8 10"7 Control -31.9667 35.7967 128 -0.89 0.3735 167 A.2 AN OVA ANALYSIS FOR IMMUNOSENSOR SENSITIVITY The Mixed Procedure Table A-6. Type 3 Tests of Fixed Effects Effec‘ NDan 13)? Veiiue Pr > F Concentration 8 54 10.32 <.0001 Time 2 54 0.34 0.7122 Concentration*Time 16 54 0.25 0.9983 Table A-7. Estimates Label Estimate Error DF Vatlue Pr > |t| 10"] vs. Control -23.8222 39.0112 54 -0.6] 0.544 10"2 vs. Control ~139.83 39.0112 54 ~3.58 0.0007 10"3 vs. Control -186.43 39.0112 54 -4.78 <.0001 10"4 vs. Control -162.7 39.0112 54 -4.17 0.0001 10"5 vs. Control -174.7 39.0112 54 -4.48 <.0001 10"6 vs. Control -191.94 39.0112 54 -4.92 <.0001 10"7 vs. Control -]88.57 39.0112 54 -4.83 <.0001 Table A-8. Least Squares Means Effect Cone. Estimate $3:de DF Vatlue Pr > |t| Cone. 10"1 233.6 27.585] 54 8.47 <.0001 Cone. 10"2 117.59 27.5851 54 4.26 <.0001 Conc. 10"3 70.9889 27.5851 54 2.57 0.0128 Conc. 10"4 94.7222 27.5851 54 3.43 0.0012 Conc. 10"S 82.7222 27.5851 54 3 0.0041 Cone. 10"6 65.4778 27.5851 54 2.37 0.0212 Conc. 10"7 68.8556 27.5851 54 2.5 0.0156 Cone. Control 257.42 27.5851 54 9.33 <.0001 168 Table A-9. Differences of Least Squares Means Effect Conc. _Conc. Estimate Standard DF t Pr > |t| Error Value Conc. 10"] 10"2 116.01 39.0112 54 2.97 0.0044 Conc. 10"] 10"3 162.61 39.0112 54 4.17 0.0001 Conc. 10"] 10"4 138.88 39.0112 54 3.56 0.0008 Conc. 10"] 10"5 150.88 39.0112 54 3.87 0.0003 Cone. 10"] 10"6 168.12 39.0112 54 4.31 <.0001 Conc. 10"] 10"7 164.74 39.01 12 54 4.22 <.0001 Conc. 10"] Control -23.8222 39.0112 54 -0.61 0.544 Cone. 10"2 10"3 46.6 39.0112 54 1.19 0.2375 Conc. 10"2 10"4 22.8667 39.0112 54 0.59 0.5602 Conc. 10"2 10"5 34.8667 39.0112 54 0.89 0.3754 Conc. 10"2 10"6 52.1111 39.0112 54 1.34 0.1872 Cone. 10"2 10"7 48.7333 39.0112 54 1.25 0.217 Conc. 10"2 Control -139.83 39.0112 54 -3.58 0.0007 Conc. 10"3 10"4 -23.7333 39.0112 54 -0.61 0.5455 Conc. 10"3 10"5 -11.7333 39.0112 54 -0.3 0.7647 Cone. 10"3 10"6 5.5111 39.0112 54 0.14 0.8882 Conc. 10"3 10"7 2.1333 39.0112 54 0.05 0.9566 Conc. 10"3 Control -186.43 39.01 12 54 -4.78 <.0001 Conc. 10"4 10"5 12 39.0112 54 0.3] 0.7596 Conc. 10"4 10"6 29.2444 39.0112 54 0.75 0.4567 Conc. 10"4 10"7 25.8667 39.0112 54 0.66 0.5101 Cone. 10"4 Control -162.7 39.0112 54 -4. 1 7 0.0001 Conc. 10"5 10"6 17.2444 39.01 12 54 0.44 0.6602 Conc. 10"5 10"7 13.8667 39.0112 54 0.36 0.7236 Conc. 10"5 Control -1 74.7 39.01 12 54 -4.48 <.0001 Cone. 10"6 10"7 -3.3778 39.0112 54 -0.09 0.9313 Conc. 10"6 Control -191.94 39.0112 54 -4.92 <.0001 Conc. 10"7 Control -188.57 39.0112 54 -4.83 <.0001 169 A.3 AN OVA ANALYSIS FOR IMMUNOSENSOR SPECIFICITY The Mixed Procedure Table A-10. Type 3 Tests of Fixed Effects Effect N151? 11))? Vine Pr > F Pathogen ] 96 2.07 0.1537 Concentration 96 0.72 0.6084 Pathogen*Concentration 96 0.83 0.5284 Table A-ll. Estimates Label Estimate Standard DF t Pr > |t| Error Value 10"] vs. Control for E. coli -50.8222 34.6163 96 -1.47 0.1453 10"2 vs. Control for E. coli 7.0167 34.6163 96 0.2 0.8398 10"3 vs. Control for E. coli -38.4889 34.6163 96 -1.11 0.269 10"4 vs. Control for E. coli -54.7333 34.6163 96 -1.58 0.1171 10"5 vs. Control for E. coli -50.7111 34.6163 96 -1.46 0.1462 10"] vs. Control for S. Enteritidis -9.4778 34.6163 96 -0.27 0.7848 10"2 vs. Control for S. Enteritidis -0.4111 34.6163 96 -0.01 0.9905 10"3 vs. Control for S. Enteritidis -l6.4333 34.6163 96 -0.47 0.6361 10"4 vs. Control for S. Enteritidis 0.9667 34.6163 96 0.03 0.9778 10"5 vs. Control for S. Enteritidis 22.3 34.6163 96 0.64 0.521 170 AA The Mixed Procedure ANOVA ANALYSIS FOR LETTUCE TESTING Table A-12. Type 3 Tests of Fixed Effects Num Den Effect DF DF F Value Pr > F Concentration 7 64 25.37 <.0001 Table A-l3. Least Squares Means Effect Estimate Standard DF t Pr > |t| Error Value Concentration 10"] 290.86 16.1587 64 18 <.0001 Concentration 10"2 123.69 16.1587 64 7.65 <.0001 Concentration 10"3 99.7889 16.1587 64 6.18 <.0001 Concentration 10"4 132.66 16.1587 64 ' 8.21 <.0001 Concentration 10"5 150.48 16.1587 64 9.31 <.0001 Concentration 10"6 99. 1 16.1587 64 6.13 <.0001 Concentration 10"7 102.37 16.1587 64 6.34 <.0001 Concentration Control 288.23 16.1587 64 17.84 <.0001 171 Table A-14. Differences of Least Squares Means Effect Conc. _Conc. Estimate Stgndard DF t Value Pr > |t| rror Conc. 10"] 10"2 167.17 22.8518 64 7.32 <.0001 C0110. 10"] 10"3 191.07 22.8518 64 8.36 <.0001 COHC- 10"] 10"4 158.2 22.8518 64 6.92 <.0001 COHC- 10"] 10"5 140.38 22.8518 64 6.14 <.0001 COHC- 10"] 10"6 191.76 22.8518 64 8.39 <.0001 CODC- 10"1 10"7 188.49 22.8518 64 8.25 <.0001 COHC- 10"] Control 2.6222 22.8518 64 0.1 1 0.909 COHC- 10"2 10"3 23.9 22.8518 64 1.05 0.2996 CODC- 10"2 10"4 -8.9667 22.8518 64 -0.39 0.6961 C0110. 10"2 10"5 -26.7889 22.8518 64 -1.17 0.2454 COHC- 10"2 10"6 24.5889 22.8518 64 1.08 0.286 Conc. 10"2 10"7 21.3222 22.8518 64 0.93 0.3543 CODC- 10"2 Control -l64.54 22.8518 64 -7.2 <.0001 COHC- 10"3 10"4 -32.8667 22.8518 64 -1.44 0.1552 C0110. 10"3 10"5 -50.6889 22.8518 64 -2.22 0.0301 COHC- 10"3 10"6 0.6889 22.8518 64 0.03 0.976 C0110. 10"3 10"7 -2.5778 22.8518 64 -0.11 0.9105 COHC- 10"3 Control -188.44 22.8518 64 -8.25 <.0001 COHC- 10"4 10"5 -17.8222 22.8518 64 -0.78 0.4383 COIN- 10"4 10"6 33.5556 22.8518 64 1.47 0.1469 COHC- 10"4 10"7 30.2889 22.8518 64 1.33 0.1897 Conc. 10"4 Control -155.58 22.8518 64 -6.81 <.0001 COHC- 10"5 10"6 51.3778 22.8518 64 2.25 0.028 COHC- 10"5 10"7 48.1111 22.8518 64 2.11 0.0392 Cone. 10"5 Control -137.76 22.8518 64 -6.03 <.0001 COHC- 10"6 10"7 -3.2667 22.8518 64 -0. 14 0.8868 COHC- 10"6 Control -189.13 22.8518 64 -8.28 <.0001 Conc. 10"7 Control —185.87 22.8518 64 -8.13 <.0001 172 A.5 ANOVA ANALYSIS FOR GROUND BEEF TESTING The Mixed Procedure Table A-15. Type 3 Tests of Fixed Effects Num Den Effect DF DF F Value Pr > F Concentration 7 64 37.93 <.0001 Table A-16. Least Squares Means Effect Conc. Estimate Standard DF t Value Pr > |t| Enor Concentration 10"] 262.02 16.0192 64 16.36 <.0001 Concentration 10"2 1 13.8 16.0192 64 7.1 <.0001 Concentration 10"3 80.2556 16.0192 64 5.0] <.0001 Concentration 10"4 75.1889 16.0192 64 4.69 <.0001 Concentration 10"5 1 14.28 16.0192 64 7.13 <.0001 Concentration 10"6 98.0889 16.0192 64 6.12 <.0001 Concentration 10"7 64.9889 16.0192 64 4.06 0.0001 Concentration Control 331.83 16.0192 64 20.71 <.0001 173 Table A-l7. Differences of Least Squares Means Effect Conc. _Conc. Estimate Stgndard DF t Value Pr > |t| rror Cone. 10"] 10"2 148.22 22.6545 64 6.54 <.0001 COHC- 10"] 10"3 181.77 22.6545 64 8.02 <.0001 CODC- 10"] 10"4 186.83 22.6545 64 8.25 <.0001 COHC- 10"1 10"5 147.74 22.6545 64 6.52 <.0001 COHC- 10"] 10"6 163.93 22.6545 64 7.24 <.0001 Cone. 10A1 10A7 197.03 22.6545 64 8.7 <.0001 CODC- 10"] Control -69.81 11 22.6545 64 -3.08 0.3030 CODC- 10"2 10"3 33 .5444 22.6545 64 1.48 0.1436 COHC- 10"2 10"4 38.6111 22.6545 64 1.7 0.0932 COHC- 10"2 10"5 -0.4778 22.6545 64 -0.02 0.9832 CODC- 10"2 10"6 15.7] 1 1 22.6545 64 0.69 0.4905 CODC- 10"2 10"7 48.8111 22.6545 64 2.15 0.035 C0110. 10"2 Control -218.03 22.6545 64 -9.62 <.0001 COHC- 10"3 10"4 5.0667 22.6545 64 0.22 0.8237 COHC- 10"3 10"5 -34.0222 22.6545 64 -1.5 0.1381 CODC- 10"3 10"6 -17.8333 22.6545 64 -0.79 0.4341 COHC- 10"3 10"7 15.2667 22.6545 64 0.67 0.5028 COHC- 10"3 Control -251.58 22.6545 64 -1 1.1 <.0001 COHC- 10"4 10"5 -39.0889 22.6545 64 -1.73 0.0893 COHC- 10"4 10"6 -22.9 22.6545 64 -1.01 0.3159 COHC- 10"4 10"7 10.2 22.6545 64 0.45 0.6541 COHC- 10"4 Control -256.64 22.6545 64 -1 1.33 <.0001 CODC- 10"5 10"6 16.1889 22.6545 64 0.71 0.4775 C0110 10"5 10"7 49.2889 22.6545 64 2.18 0.0333 COHC- 10"5 Control -2 1 7 .56 22.6545 64 -9.6 <.0001 CODC- 10"6 10"7 33.] 22.6545 64 1.46 0.1489 CODC- 10"6 Control -233.74 22.6545 64 -10.32 <.0001 COHC- 10"7 Control -266.84 22.6545 64 -] 1.78 <.0001 174 A.6 ANOVA ANALYSIS FOR WHOLE MILK TESTING The Mixed Procedure Table A-18. Type 3 Tests of Fixed Effects Num Den DF DF Concentration 7 64 32.15 <.0001 Effect F Value Pr > F Table A-l9. Least Squares Means Effect Conc. Estimate Standard DF tValue Pr > It] Enor Concentration 10"] 267.94 18.9017 64 14.18 <.0001 Concentration 10"2 274.57 18.9017 64 14.53 <.0001 Concentration 10"3 1 16.21 18.9017 64 6.15 <.0001 Concentration 10"4 103.44 18.9017 64 5 .47 <.0001 Concentration 10"5 116.61 18.9017 64 6.17 <.0001 Concentration 10"6 84.6889 18.9017 64 4.48 <.0001 Concentration 10"7 76.4 1 8.901 7 64 4.04 0.0001 Concentration Control 353.88 18.9017 64 18.72 <.0001 175 Table A-20. Differences of Least Squares Means Standard Y 1 Effect Conc. _Conc. Estimate Error DF tValue ‘ Pr>tl Cone. 1041 1042 -6.6222 26.731 64 -025 ' 0.8051 Conc- 1041 1043 151.73 26.731 64 5.68 <.0001 - Cone. 1041 1044 164.5 26.731 64 6.15 <.0001 ‘ Cone. 1041 1045 151.33 26.73] 64 5.66 <.0001 Cone. 1041 1046 183.26 26.73] 64 6.86 <.0001 Conc- 1041 1047 191.54 26.731 64 7.17 <.0001 Cone. 1041 Control -85.9333 26.731 64 -321 0.2020 Conc- 1042 1043 158.36 26.731 64 5.92 <.0001 Cone. 1042 1044 171.12 26.73] 64 6.4 <.0001 Cone. 1042 1045 157.96 26.731 64 5.91 <.0001 Cone. 1042 1046 189.88 26.731 64 7.1 <.0001 Cone. 1042 1047 198.17 26.731 64 7.41 <.0001 Cone. 1042 Control -7931 1 1 26.731 64 -297 0.4002 Cone. 1043 1044 12.7667 26.731 64 0.48 0.6346 Cone. 1043 1045 .04 26.731 64 -001 0.9881 Cone. 1043 10"6 31.5222 26.731 64 1.18 0.2427 Cone. 1043 1047 39.811] 26.731 64 1.49 0.1413 Cone. 1043 Control -237.67 26.73] 64 -8.89 <.0001 Cone. 1044 1045 -13.1667 26.731 64 -049 0.624 Cone. 1044 1046 18.7556 26.731 64 0.7 0.4854 Conc- 1044 1047 27.0444 26.731 64 1.01 0.3155 Cone. 1044 Control -25043 26.73] 64 -937 <.0001 Cone. 1045 1046 31.9222 26.731 64 1.19 0.2368 Cone. 1045 1047 40.21 1 1 26.731 64 1.5 0.1374 Cone. 1045 Control -23727 26.731 64 -8.88 <.0001 Cone. 1046 1047 8.2889 26.731 64 0.31 0.7575 Cone. 1046 Control -269.19 26.731 64 -1007 <.0001 Cone. 1047 Control -277.48 26.73] 64 -10.38 <.0001 176 A.7 The Mixed Procedure Table A-21. Type 3 Tests of Fixed Effects ANOVA ANALYSIS FOR SPCE BIOSENSOR SENSITIVITY Effec‘ NDan 11):}:1 Veilue P‘ > F DNA Cone. 5 12 72.19 <.0001 Table A-22. Estimates Label Estimate Standard DF t Value Pr > |t| Error 0.00] vs. Control -0.8033 1.9949 12 -0.4 0.6943 0.01 vs. Control 14.3433 1.5827 12 9.06 <.0001 0.1 vs. Control 19.9767 2.2137 12 9.02 <.0001 1 vs. Control 37.3767 2.7075 12 13.8 <.0001 10 vs. Control 330.21 143.02 12 2.31 0.0396 Table A-23. Least Squares Means Effect DNA Conc. Estimate Standard DF t Pr > |t| Error Value DNA Conc. 0.001 4.72 1.2258 12 3.85 0.0023 DNA Conc. 0.01 19.8667 0.1667 12 119.2 <.0001 DNA Conc. 0.1 25.5 1.5567 12 16.38 <.0001 DNA Conc. 1 42.9 2.203 12 19.47 <.0001 DNA Conc. 10 335.73 143.01 12 2.35 0.0369 DNA Conc. Control 5.5233 1.5739 12 3.51 0.0043 177 Table A-24. Differences of Least Squares Means DNA DNA Standard t Effect Conc. _Conc. Estimate Error DF Value Pr > m DNA Conc. 0.001 0.01 -15.1467 1.237 ] 12 -12.24 <.0001 DNA Conc. 0.001 0.1 -20.78 1.9814 12 -10.49 <.0001 DNA Conc. 0.001 1 -38.18 2.5211 12 -15.14 <.0001 DNA Cone. 0.001 10 -331.01 143.02 12 -2.31 0.0392 DNA Cone. 0.001 Control -0.8033 1.9949 12 -0.4 0.6943 DNA Conc. 0.01 0.1 —5.6333 1.5656 12 -3.6 0.0037 DNA Cone. 0.01 1 -23.0333 2.2093 12 -10.43 <.0001 DNA Conc. 0.01 10 -315.87 143.01 12 -2.21 0.0474 DNA Conc. 0.0] Control 14.3433 1.5827 12 9.06 <.0001 DNA Cone. 0.1 1 -17.4 2.6975 12 -6.45 <.0001 DNA Conc. 0.1 10 -310.23 143.02 12 -2.17 0.0509 DNA Cone. 0.1 Control 19.9767 2.2137 12 9.02 <.0001 DNA Conc. 1 10 -292.83 143.03 12 -2.05 0.0432 DNA Conc. 1 Control 37.3767 2.7075 12 13.8 <.0001 DNA Cone. 10 Control 330.21 143.02 12 2.31 0.0396 178 APPENDIX B: DATA Figure B-l. SEM images of (A) Iron oxide NPs and (B) EAPM NPs 179 N O O 2 mglml g 150 O '35 3 100 - C .‘3 .2 8 50 - I! 0 _ Control 10" 7 10" 4 10" 1 Cell Concentration (CFUImI) 100 20 mglml g 75 — O i‘, 8 50 - t: 3 fl 8 25 — a: 0 _ Control 10" 7 10" 4 10" 1 Cell Concentration (CFUImI) 200 100 mglml 150 - Resistance (kOhm) 01 8 O O Control 10" 7 10" 4 10" 1 Cell Concentration (CFUImI) Figure B-2. Immunosensor responses of different EAPM concentrations in B. cereus 180 N O O 100 pglml .15: 150 — o i‘. 8 c 100 3 .2 § 50 4 o 2 Control 10"7 10"4 10"1 Cell Concentration (CFU/ml) 100 150 pglml ’E‘ .C O 5 8 50 — C 3 .2 In 0 n: 0 _ Control 10A 7 10" 4 10" 1 Cell Concentration (CFU/ml) 250 pglml A 300 4 E S 2 V 200 — Q) U C 8 “.7, 100 d) n: o _ Control 10"7 10"4 10” Cell Concentration (CFUIml) Figure B-3. Immunosensor responses of different B. cereus antibody concentrations. 181 B. anthracis spores monitored after 10 day incubation period Figure B-4. Light microscopy images of B. anthracis spores after two different incubation periods. (Final Spore Count — 4.2 X 108 spores/ml) 182 REFERENCES Abrahamsson, D., Kriz, K., Lu, M., and Kriz, D. (2004) A preliminary study on DNA detection based on relative magnetic permeability measurements and histone H1 conjugated superparamagnetic nanoparticles as magnetic tracers. Biosensors and Bioelectronics 19(11): 1549-1557. Aguilar, Z.P., and Sirisena, M. (2007) Development of automated amperometric detection of antibodies against Bacillus anthracis protective antigen. Analytical and Bioanalytical Chemistry 389(2): 507-515. Ahuja, T., Mir, I.A., Kumar, D., and Rajesh (2007) Biomolecular immobilization on conducting polymers for biosensing applications. Biomaterials 28(5): 791-805. Alam, J., Riaz, U., and Ahmad, S. (2007) Effect of ferrofluid concentration on electrical and magnetic properties of the Fe3O4/PANI nanocomposites. Journal of Magnetism and Magnetic Materials 314(2): 93-99. Arora, K., Prabhakar, N., Chand, S., and Malhotra, B.D. (2007) Escherichia coli genosensor based on polyaniline. Analytical Chemistry 79(16): 6152-6158. Asmatulu, R., Zalich, M.A., Claus, R., and Riffle, J .S. (2005) Synthesis, characterization and targeting of biodegradable magnetic nanocomposite particles by external magnetic fields. Journal of Magnetism and Magnetic Materials 292: 108-119. Bard, Al, and Faulkner, LR. (2001) Electrochemical Methods: Fundamentals and Applications John Wiley & Sons, Inc. Bereket, G. and Sahin, Y. (2005) Electrochemical synthesis and anti-corrosive properties of polyaniline, poly(2-anisidine), and poly(ani1ine-co-2-anisidine) films on stainless steel. Progress in Organic Coatings 54(1): 63-72. Berkenpas, E., Millard, P., and da Cunha, MP. (2006) Detection of Escherichia coli 0157 : H7 with langasite pure shear horizontal surface acoustic wave sensors. Biosensors & Bioelectronics 21(12): 2255-2262. 183 Bisoffi, M., Hjelle, B., Brown, D.C., Branch, D.W., Edwards, T.L., Brozik, S.M. et al. (2008) Detection of viral bioagents using a shear horizontal surface acoustic wave biosensor. Biosensors & Bioelectronics 23(9): 1397-1403. Boissiere, M., Meadows, P.J., Brayner, R., Helary, C., Livage, J ., and Coradin, T. (2006) Turning biopolymer particles into hybrid capsules: the example of silica/alginate nanocomposites. Journal of Materials Chemistry 16(12): 1178-1182. Bothara, M., Venkatrarnan, V., Reddy, R.K.K., Barrett, T., Carruthers, J., and Prasad, S. (2008) Nanomonitors: electrical immunoassays for protein biomarker profiling. Nanomedicine 3(4): 423-436. Branch, D.W., and Brozik, S.M. (2004) Low-level detection of a Bacillus anthracis simulant using Love-wave biosensors on 36 degrees YX LiTaO3. Biosensors & Bioelectronics 19(8): 849-859. Buijs, J ., vandenBerg, P.A.W., Lichtenbelt, J .W.T., Norde, W., and Lyklema, J. (1996) Adsorption dynamics of IgG and its F(ab')(2) and Fe fragments studied by reflectometry. Journal of Colloid and Interface Science 178(2): 594-605. Campbell, GA, and Mutharasan, R. (2006) Piezoelectric-excited millimeter-sized cantilever (PEMC) sensors detect Bacillus anthracis at 300 spores/mL. Biosensors & Bioelectronics 21(9): 1684-1692. Carrascosa, L.G., Moreno, M., Alvarez, M., and Lechuga, L.M. (2006) Nanomechanical biosensors: a new sensing tool. T rac-Trends in Analytical Chemistry 25(3): 196-206. CDC (2001) Anthrax Overview [WWW document]. URL, http://www.bt.cdc.goWaggent/anthrax/SlideSetAnthrax.pdf, Accessed March 15, 2009. CDC (2009) Bioterrorism Agents/Diseases [WWW document]. URL, http://www.bt.cdc.gov/agent/agentlist—categorvasp, Accessed March 15, 2009. Chang, H.X., Yuan, Y., Shi, N.L., and Guan, Y.F. (2007a) Electrochemical DNA biosensor based on conducting polyaniline nanotube array. Analytical Chemistry 79(13): 5111-5115. 184 Chang, H.X., Yuan, Y., Shi, N.L., and Guan, Y.F. (2007b) Electrochemical DNA biosensor based on conducting polyaniline nanotube array. Analytical Chemistry 79(13): 5111-5115. Chang, M., Glynn, M.K., and Groseclose, S.L. (2003) Endemic, notifiable bioterrorism related diseases, United States, 1002-1999. Emerging Infectious Diseases 9(5): 556-564. Chang, Z., Pan, H., Zhao, K., Chen, M., He, P.G., and Fang, Y.Z. (2008) Electrochemical DNA biosensors based on palladium nanoparticles combined with carbon nanotubes. Electroanalysis 20(2): 131-136. Chang, Z., Zhou, J., Zhao, K., Zhu, N., He, P., and Fang, Y. (2006) Ru(bpy)32+-doped silica nanoparticle DNA probe for the electrogenerated chemiluminescence detection of DNA hybridization. Electrochimica Acta 52(2): 575-580. Chen, L.L., Deng, L., Liu, L.L., and Peng, Z.H. (2007) Immunomagnetic separation and MS/SPR end-detection combined procedure for rapid detection of Staphylococcus aureus and protein A. Biosensors & Bioelectronics 22(7): 1487-1492. Cheun, H.I., Makino, S.I., Wataral, M., Shirahata, T., Uchida, I., and Takeshi, K. (2001) A simple and sensitive detection system for Bacillus anthracis in meat and tissue. Journal of Applied Microbiology 91(3): 421-426. Chung, J.W., Kim, S.D., Bernhardt, R., and Pyun, J.C. (2005) Application of SPR biosensor for medical diagnostics of human hepatitis B virus (hHBV). Sensors and Actuators B—Chemical 11 1: 416-422. Cieslak, T.J., and Eitzen, EM. (1999) Clinical and epidemiological principles of anthrax. Emerging Infectious Diseases 5: 552-555. Collier, R.J., and Young, J.A.T. (2003) Anthrax toxin. Annual Review of Cell and Developmental Biology 19: 45-70. Comparelli,R., Curri,M.L., Cozzoli,P.D., and Striccoli,M. (2007) Optical Biosensing Based on Metal and Semiconductor Colloidal Nanocrystals. In Nanomaterials for Biosensors. Kumar,C.S.S.R. (ed). Wiley-VCH, pp. 123-174. 185 Cooper, M.A., and Singleton, VT (2007) A survey of the 2001 to 2005 quartz crystal microbalance biosensor literature: applications of acoustic physics to the analysis of biomolecular interactions. Journal of Molecular Recognition 20(3): 154-184. Costa-Femandez, J .M. (2006) Optical sensors based on luminescent quantum dot. Analytical and Bioanalytical Chemistry 384(1): 37-40. Cui, Y., Zhong, Z.H., Wang, D.L., Wang, W.U., and Lieber, CM. (2003) High performance silicon nanowire field effect transistors. Nano Letters 3(2): 149-152. Dai, Z., Bai, H., Hong, M., Zhu, Y., Bao, J ., and Shen, J. (2008) A novel nitrite biosensor based on the direct electron transfer of hemoglobin immobilized on CdS hollow nanospheres. Biosensors and Bioelectronics 23(12): 1869-1873. Dallas, P., Moutis, N., Devlin, E., Niarchos, D., and Petridis, D. (2006) Characterization, electrical and magnetic properties of polyaniline/maghemite nanocomposites. Nanotechnology 17(19): 5019-5026. Dastagir, T., Forzani, E.S., Zhang, R., Amlani, I., Nagahara, L.A., Tsui, R., and Tao, N. (2007) Electrical detection of hepatitis C virus RNA on single wall carbon nanotube-field effect transistors. Analyst 132(8): 73 8-740. Davila, A.P., Jang, J., Gupta, A.K., Walter, T., Aronson, A., and Bashir, R. (2007) Microresonator mass sensors for detection of Bacillus anthracis Sterne spores in air and water. Biosensors & Bioelectronics 22(12): 3028-3035. Delvaux, M., Duchet, J ., Stavaux, P.Y., Legras, R., and moustier-Champagne, S. (2000) Chemical and electrochemical synthesis of polyaniline micro- and nano-tubules. Synthetic Metals 113(3): 275-280. Deng, J., He, C., Peng, Y., Wang, J., Long, X., Li, R, and Chan, A.S.C. (2003) Magnetic and conductive F e304-polyaniline nanoparticles with core-shell structure. Synthetic Metals 139(2): 295-301. Deng, Z., and Alocilja, EC. (2008) Characterization of Nanoporous Silicon-Based DNA Biosensor for the Detection of Salmonella Enteritidis. Sensors Journal, IEEE 8(6): 775- 780. 186 Deobagkar, D.D., Limaye, V., Sinha, S., and Yadava, R.D.S. (2005) Acoustic wave immunosensing of Escherichia coli in water. Sensors and Actuators B-Chemical 104(1): 85-89. Desai, P.T., Walsh, M.K., and Weirner, BC. (2008). Solid-phase capture of pathogenic bacteria by using gangliosides and detection with real-time PCR. Applied and Environmental Microbiology, 74(7): 2254-2258. Diamond, D. (1998) Principles of Chemical and Biological Sensors John Wiley & Sons, Inc. Ding, H., Liu, X.M., Wan, M., and Fu, S.Y. (2008) Electromagnetic functionalized cage- like polyaniline composite nanostructures. Journal of Physical Chemistry B 112(31): 9289-9294. Dong, X.C., Fu, D.L., Xu, Y.P., Wei, J.Q., Shi, Y.M., Chen, P., and Li, L]. (2008) Label- free electronic detection of DNA using simple double-walled carbon nanotube resistors. Journal of Physical Chemistry C 112(26): 9891-9895. Drummond, T.G., Hill, M.G., and Barton, J.K. (2003) Electrochemical DNA sensors. Nature Biotechnology 21(10): 1192-1199. Duic, L., Mandic, Z., and Kovacicek, F. (1994) Effect of supporting electrolyte on the electrochemical synthesis, morphology, and conductivity of polyaniline. Journal of Polymer Science, Part A: Polymer Chemistry 32(1): 105-111. Duic, L., Mandic, Z., and Kovac, S. (1995) Polymer-dimer distribution in the electrochemical synthesis of polyaniline. Electrochimica Acta 40(11): 1681-1688. Dungchai, W., Siangproh, W., Chaicumpa, W., Tongtawe, P., and Chailapakul, O. (2008) Salmonella typhi determination using voltammetric amplification of nanoparticles: A highly sensitive strategy for metalloimmunoassay based on a copper-enhanced gold label. T alanta 77(2): 727-732. Edelstein, R.L., Tarnanaha, C.R., Sheehan, P.E., Miller, M.M., Baselt, D.R., Whitman, L.J., and Colton, RI. (2000) The BARC biosensor applied to the detection of biological warfare agents. Biosensors and Bioelectronics 14(10-11): 805-813. Eggins, BR. (2002) Chemical Sensors and Biosensors John Wiley & Sons, Ltd. 187 El-Tantawy, F., bdel-Kader, K.M., Kaneko, F., and Sung, Y.K. (2004) Physical properties of CdS-poly (vinyl alcohol) nanoconducting composite synthesized by organosol techniques and novel application potential. European Polymer Journal 40(2): 415-430. Elsholz, B., Worl, R., Blohm, L., Albers, J., Feucht, H., Grunwald, T. et al. (2006) Automated Detection and Quantitation of Bacterial RNA by Using Electrical Microarrays. Analytical Chemistry 78(14): 4794-4802. Epstein, A.J., Ginder, J .M., Zuo, F., Woo, H.S., Tanner, D.B., Richter, A.F. et al. (1987) Insulator-To-Metal Transition in Polyaniline - Effect of Protonation in Emeraldine. Synthetic Metals 21(1): 63-70. Epstein, J .R., Biran, I., and Walt, DR. (2002) F luorescence-based nucleic acid detection and microarrays. Analytica Chimica Acta 469(1): 3-36. Ercole, C., Del Gallo, M., Mosiello, L., Baccella, S., and Lepidi, A. (2003) Escherichia coli detection in vegetable food by a potentiometric biosensor. Sensors and Actuators B- Chemical 91(1-3): 163-168. Erickson, D., Manda], S., Yang, A.H.J., and Cordovez, B. (2008) Nanobiosensors: optofluidic, electrical and mechanical approaches to biomolecular detection at the nanoscale. Microfluidics and Nanofluidics 4(1-2): 33-52. Erickson, M.C., and Kornacki, IL. (2003) Bacillus anthracis: Current knowledge in relation to contamination of food. Journal of Food Protection 66(4): 691-699. Fan, Y., Chen, X.T., Trigg, A.D., Tung, C.H., Kong, J .M., and Gao, Z.Q. (2007) Detection of microRNAs using target-guided formation of conducting polymer nanowires in nanogaps. Journal of the American Chemical Society 129(17): 5437-5443. Farabullini, F., Lucarelli, F., Palchetti, I., Marrazza, G., and Mascini, M. (2007) Disposable electrochemical genosensor for the simultaneous analysis of different bacterial food contaminants. Biosensors & Bioelectronics 22(7): 1544-1549. Feast, W.J., Tsibouklis, J., Pouwer, K.L., Groenendaal, L., and Meijer, E.W. (1996) Synthesis, processing and material properties of conjugated polymers. Polymer 37(22): 5017-5047. 188 Fu, J., Park, B., Siragusa, G., Jones, L., Tripp, R., Zhao, Y.P., and Cho, Y.J. (2008) An Au/Si hetero-nanorod-based biosensor for Salmonella detection. Nanotechnology 19(15). Galipeau, D.W., Story, P.R., Vetelino, K.A., and Mileham, RD. (1997) Surface acoustic wave microsensors and applications. Smart Materials & Structures 6(6): 658-667. Gao, Z.Q., Agarwal, A., Trigg, A.D., Singh, N., Fang, C., Tung, C.H. et al. (2007) Silicon nanowire arrays for label-free detection of DNA. Analytical Chemistry 79(9): 3291-3297. Geng, T., Uknalis, J ., Tu, S.I., and Bhunia, AK. (2006) Fiber-optic biosensor employing Alexa-Fluor conjugated antibody for detection of Escherichia coli 0157: H7 from ground beef in four hours. Sensors 6(8): 796-807. Gerard, M., Chaubey, A., and Malhotra, B.D. (2002) Application of conducting polymers to biosensors. Biosensors & Bioelectronics 17(5): 345-359. Goldman, E.R., Clapp, A.R., Anderson, G.P., Uyeda, H.T., Mauro, J.M., Medintz, IL, and Mattoussi, H. (2004) Multiplexed toxin analysis using four colors of quantum dot fluororeagents. Analytical Chemistry 76(3): 684-688. Gong, P., He, X., Wang, K., Tan, W., Xie, W., Wu, P., and Li, H. (2008) Combination of functionalized nanoparticles and polymerase chain reaction-based method for SARS-CoV gene detection. Journal of Nanoscience and Nanotechnology 8(1): 293-300. Gorman,C.B., and Grubbs, RH. (1991) Conjugated Polymers: The Interplay Between Synthesis, Structure, and Properties. In Conjugated Polymers. Bredas, J .L., and Silbey, R. (eds). Kluwer Academic Publishers, pp. 1-48. Gospodinova, N., Mokreva, P., and Terlemezyan, L. (1996) Concomitant processes in the redox switching of polyaniline. Polymer International 41(1): 79-84. Grzeszczuk, M., and Szostak, R. (2003) Electrochemical and Raman studies on the redox switching hysteresis of polyaniline. Solid State Ionics 157(1-4): 257-262. Hafeman, D.G., Parce, J .W., and Mcconell,H.M. (1988) Light-addressable potentiometric sensor for biochemical systems. Science 240(4856): 1182-1185. 189 Hahm, J ., and Lieber, CM. (2004) Direct ultrasensitive electrical detection of DNA and DNA sequence variations using nanowire nanosensors. Nano Letters 4(1): 51-54. Hansen, J .A., Wang, J., Kawde, A.N., Xiang, Y., Gothelf, K.V., and Collins, G. (2006) Quantum-dot/aptamer-based ultrasensitive multi-analyte electrochemical biosensor. Journal of the American Chemical Society 128(7): 2228-2229. Heeger, A., and Smith, P. (1991) Solution Processing of Conducting Polymers: Opportunities for Science and Technology. In Conjugated Polymers. Bredas, J.L., and Silbey, R. (eds). Kluwer Academic Publishers, pp. 141-210. Hill, H.D., Vega, RA, and Mirkin, CA. (2007) Nonenzymatic detection of bacterial genomic DNA using the bio bar code assay. Analytical Chemistry 79(23): 9218-9223. Hnaiein, M., Hassen, W.M., Abdelghani, A., Fournier-Wirth, C., Coste, J ., Bessueille, F. et al. (2008) A conductometric immunosensor based on firnctionalized magnetite nanoparticles for E. coli detection. Electrochemistry Communications 10(8): 1152-1154. Hoa, X.D., Kirk, AG, and Tabrizian, M. (2007) Towards integrated and sensitive surface plasmon resonance biosensors: A review of recent progress. Biosensors & Bioelectronics 23: 151-160. Hohnholz, D., Okuzaki, H., and Macdiarmid, AG. (2005) Plastic electronic devices through line patterning of conducting polymers. Advanced Functional Materials 15(1): 51-56. Homola, J. (2008) Surface plasmon resonance sensors for detection of chemical and biological species. Chemical Reviews 108(2): 462-493. Hong, S.Y., and Park, S.M. (2005) Electrochemistry of conductive polymers 36. pH dependence of polyaniline conductivities studied by current-sensing atomic force microscopy. Journal of Physical Chemistry B 109(19): 93 05-93 1 0. llic, B., Czaplewski, D., Zalalutdinov, M., Craighead, H.G., Neuzil, P., Campagnolo, C., and Batt, C. (2001) Single cell detection with micromechanical oscillators. Journal of Vacuum Science & Technology B 19(6): 2825-2828. Inglesby, T.V. (2000) Anthrax as a biological weapon: Medical and public health management. Jama-Journal of the American Medical Association 283(15): 1963. 190 Inglesby, T.V., Dennis, D.T., Henderson, D.A., Bartlett, J.G., Ascher, M.S., Eitzen, E. et al. (2000) Plague as a biological weapon - Medical and public health management. Jama- Journal of the American Medical Association 283(17): 2281-2290. Jemigan, J.A., Stephens, D.S., Ashford, D.A., Omenaca, C., Topiel, M.S., Galbraith, M. et al. (2001) Bioterrorism-related inhalational anthrax: The first 10 cases reported in the United States. Emerging Infectious Diseases 7(6): 933-944. Jiang, J., Li, LC, and Xu, F. (2006) Preparation, characterization and magnetic properties of PANI/La-substituted LiNi ferrite nanocomposites. Chinese Journal of Chemistry 24(12): 1804-1809. Jin,X., Gao,Z., Pan,H., Zhu,H., Zhou,M., and Chen,H. The surface acoustic wave biosensor for detecting the gene of Staphylococal Enterotoxin B. Proceedings of the Intemationa] Symposium on Test and Measurement 1, 261-264. 2003. Johnson, L., Gupta, A.T.K., Ghafoor, A., Akin, D., and Bashir, R. (2006) Characterization of vaccinia virus particles using microscale silicon cantilever resonators and atomic force microscopy. Sensors and Actuators B-Chemical 115(1): 189-197. Jyoung, J.Y., Hong, SH, Lee, W., and Choi, J.W. (2006) Immunosensor for the detection of Vibrio cholerae 01 using surface plasmon resonance. Biosensors & Bioelectronics 21(12): 2315-2319. Kaittanis, C., Naser, SA, and Perez, J .M. (2007) One-step, nanoparticle-mediated bacterial detection with magnetic relaxation. Nano Letters 7(2): 380-383. Kaneto, K., Fujisue, H., Kunifusa, M., and Takashima, W. (2007) Conducting polymer soft actuators based on polypyrrole films - energy conversion efficiency. Smart Materials & Structures 16(2): 8250-8255. Katz, E., and Willner, I. (2003) Probing biomolecular interactions at conductive and semiconductive surfaces by impedance spectroscopy: Routes to impedimetric immunosensors, DNA-Sensors, and enzyme biosensors. Electroanalysis 15(11): 913-947. Kaufmann, A.F., Meltzer, M.l., and Schmid, GP. (1997) The eonomic impact of a bioterrotism attack: Are prevention and postattack intervention programs justifiable? Emerging Infectious Diseases 3(2): 83-94. 191 Kim, H., Kane, M.D., Kim, S., Dominguez, W., Applegate, B.M., and Savikhin, S. (2007a) A molecular beacon DNA microarray system for rapid detection of E. coli 0157: H7 that eliminates the risk of a false negative signal. Biosensors & Bioelectronics 22(6): 1041-1047. Kim, J.H., Cho, J.H., Cha, G.S., Lee, C.W., Kim, H.B., and Pack, SH. (2000) Conductimetric membrane strip immunosensor with polyaniline-bound gold colloids as signal generator. Biosensors & Bioelectronics 14(12): 907-915. Kim, S.N., Rusling, J.F., and Papadimitrakopoulos, F. (2007b) Carbon nanotubes for electronic and electrochemical detection of biomolecules. Advanced Materials 19(20): 3214-3228. Ko, S., and Jang, J. (2008) Label-free target DNA recognition using oligonucleotide- functionalized polypyrrole nanotubes. Ultramicroscopy 108(10): 1328-1333. Ko, SH, and Grant, SA. (2006) A novel FRET-based optical fiber biosensor for rapid detection of Salmonella Typhimurium. Biosensors & Bioelectronics 21(7): 1283-1290. Kong, J ., Franklin, N.R., Zhou, C.W., Chapline, M.G., Peng, S., Cho, K.J., and Dai, H]. (2000) Nanotube molecular wires as chemical sensors. Science 287(5453): 622-625. Krebs, M.D., Mansfield, B., Yip, R, Cohen, S.J., Sonenshein, A.L., Hitt, BA, and Davis, CE. (2006) Novel technology for rapid species-specific detection of Bacillus spores. Biomolecular Engineering 23(2-3): 119-127. Kriz, K., Gehrke, J ., and Kriz, D. (1998) Advancements toward magneto immunoassays. Biosensors and Bioelectronics 13(7-8): 817-823. Kryszewski, M., and J eszka, J .K. (1998) Nanostructured conducting polymer composites - superparamagnetic particles in conducting polymers. Synthetic Metals 94(1): 99-104. Laird, P.W., Zijderveld, A., Linders, K., Rudnicki, M.A., Jaenisch, R., and Bems, A. (1991) Simplified Mammalian DNA Isolation Procedure. Nucleic Acids Research 19(15): 4293. Lan, Y.B., Wang, S.Z., Yin, Y.G., Hoffmann, W.C., and Zheng, X.Z. (2008) Using a Surface Plasmon Resonance Biosensor for Rapid Detection of Salmonella Typhimurium in Chicken Carcass. Journal of Bionic Engineering 5(3): 23 9—246. 192 Lange, K., Rapp, BE, and Rapp, M. (2008) Surface acoustic wave biosensors: a review. Analytical and Bioanalytical Chemistry 391(5): 1509-1519. Langer, J .J ., Filipiak, M., Kecinska, J ., Jasnowska, J ., Wodarczak, J ., and Buadowski, B. Polyaniline biosensor for choline determination. Proceedings of the 9th International Fischer Symposium, July 21, 2003 - July 23, 2003. 573(1), 140-145. 2004. Munich, Germany, Elsevier. Surface Science. Lazcka, 0., Del Carnpo, P.J., and Munoz, F.X. (2007) Pathogen detection: A perspective of traditional methods and biosensors. Biosensors & Bioelectronics 22(7): 1205-1217. Lee, H., Lee, E., Kim, D.K., Jang, N.K., Jeong, Y.Y., and Jon, S. (2006) Antibiofouling polymer-coated superparamagnetic iron oxide nanoparticles as potential magnetic resonance contrast agents for in vivo cancer imaging. Journal of the American Chemical Society 128(22): 7383-7389. Lee, SH, Stubbs, D.D., Caimey, J ., and Hunt, W.D. (2005) Rapid detection of bacterial spores using a quartz crystal microbalance (QCM) immunoassay. Ieee Sensors Journal 5(4): 737-743. Leppla, SH. (1982) Anthrax Toxin Edema Factor - A Bacterial Adenylate-Cyclase That Increases Cyclic-Amp Concentrations in Eukaryotic Cells. Proceedings of the National Academy of Sciences of the United States of America-Biological Sciences 79(10): 3162- 3166. Lermo, A., Campoy, S., Barbe, J., Hernandez, S., Alegret, S., and Pividori, M. (2007) In situ DNA amplification with magnetic primers for the electrochemical detection of food pathogens. Biosensors & Bioelectronics 22(9-10): 2010-2017. Li, C., Liu, Y., Li, L., Du, Z., Xu, S., Zhang, M. et al. (2008) A novel amperometric biosensor based on NiO hollow nanospheres for biosensing glucose. Talanta 77(1): 455- 459. Li, G., Jiang, L., and Peng, H. (2007a) One-dimensional polyaniline nanostructures with controllable surfaces and diameters using vanadic acid as the oxidant. Macromolecules 40(22): 7890-7894. Li, G., Yan, S., Zhou, E., and Chen, Y. (2006) Preparation of magnetic and conductive NiZn ferrite-polyaniline nanocomposites with core-shell structure. Colloids and Surfaces A: Physicochemical and Engineering Aspects 276(1-3): 40-44. 193 Li, X., Shen, J .Y., Wan, M.X., Chen, Z.J., and Wei, Y. (2007b) Core-shell structured and electro-magnetic functionalized polyaniline composites. Synthetic Metals 157(13-15): 575-579. Li, Y.G., Cu, Y.T.H., and Luo, D. (2005) Multiplexed detection of pathogen DNA with DNA-based fluorescence nanobarcodes. Nature Biotechnology 23(7): 885-889. Liu, Y.S., Walter, T.M., Chang, W.J., Lim, K.S., Yang, L.J., Lee, S.W. et al. (2007) Electrical detection of germination of viable model Bacillus anthracis spores in microfluidic biochips. Lab on A Chip 7(5): 603-610. Liu, Y.J., Yao, D.J., Chang, H.Y., Liu, C.M., and Chen, C. (2008) Magnetic bead-based DNA detection with multi-layers quantum dots labeling for rapid detection of Escherichia coli 01 57:H7. Biosensors and Bioelectronics 24(4): 558-565. Lizarraga, L., Andrade, B.M., and Molina, F.V. (2004) Swelling and volume changes of polyaniline upon redox switching. Journal of Electroanalytical Chemistry 561(1-2): 127- 135. Ludwig, F., Heim, E., Menzel, D., and Schilling, M. (2006) Investigation of superparamagnetic Fe304 nanoparticles by fluxgate magnetorelaxometry for use in magnetic relaxation immunoassays. Journal of Applied Physics 99(8). Luo, X.L., Morrin, A., Killard, A.J., and Smyth, MR. (2006) Application of nanoparticles in electrochemical sensors and biosensors. Electroanalysis 18(4): 319-326. Lv, R., Zhang, S., Shi, Q., and Kan, J. (2005) Electrochemical synthesis of polyaniline nanoparticles in the presence of magnetic field and erbium chloride. Synthetic Metals 150(2): 115-122. Lyons, M.E.G. (1994) Electroactive Polymer Electrochemistry Plenum Press. Macdiarmid, A.G., Chiang, J .C., Richter, AF, and Epstein, A.J. (1987) Polyaniline - A New Concept in Conducting Polymers. Synthetic Metals 18(1-3): 285-290. Mairal, T., Ozalp, V.C., Sanchez, P.L., Mir, M., Katakis, I., and O'Sullivan, CK. (2008) Aptamers: molecular tools for analytical applications. Analytical and Bioanalytical Chemistry 390(4): 989-1007. 194 Malhotra, B.D., Chaubey, A., and Singh, SP. (2006) Prospects of conducting polymers in biosensors. Analytica Chimica Acta 578(1): 59-74. Mao, X.L., Yang, L.J., Su, XL, and Li, Y.B. (2006) A nanoparticle amplification based quartz crystal microbalance DNA sensor for detection of Escherichia coli 0157:H7. Biosensors & Bioelectronics 21(7): 1178-1185. Marinakos, S.M., Chen, SH, and Chilkoti, A. (2007) Plasmonic detection of a model analyte in serum by a gold nanorod sensor. Analytical Chemistry 79(14): 5278-5283. Marquette, CA, and Blum, L.J. (2006) State of the art and recent advances in immunoanalytical systems. Biosensors & Bioelectronics 21(8): 1424- 1 43 3. Mason, J.T., Xu, L., Sheng, Z., and O'Leary, T.J. (2006) A lioposome-PCR assay for the ultrasensitive detection of biological toxins. Nature Biotechnology 24(5): 555-557 Mathew, F .P., and Alocilja, EC. (2005) Porous silicon-based biosensor for pathogen detection. Biosensors and Bioelectronics 20(8 SPEC. ISS.): 1656-1661. Mazur, M., Krysinski, P., and Palys, B. (2002) Preparation of ultrathin films of polyaniline and its derivatives by electrochemical deposition on thiol modified gold. Journal of Electroanalytical Chemistry 533(1-2): 145-152. Mehlen, A., Goeldner, M., Ried, S., Stindl, S., Ludwig, W., and Schleifer, K.H. (2004) Development of a fast DNA-DNA hybridization method based on melting profiles in microplates. Systematic and Applied Microbiology 27(6): 689-695. Mishra, N.N., Maki, W.C., Cameron, E., Nelson, R., Winterrowd, P., Rastogi, S.K. et al. (2008) Ultra-sensitive detection of bacterial toxin with silicon nanowire transistor. Lab on A Chip 8(6): 868-871. Mock, M., and Fouet, A. (2001) Anthrax. Annual Review of Microbiology 55: 647-671. Mock, M., and Mignot, T. (2003) Anthrax toxins and the host: a story of intimacy. Cellular Microbiology 5(1): 15-23. 195 Moll, N., Pascal, E., Dinh, D.H., Lachaud, J.L., Vellutini, L., Pillot, JP. et al. (2008) Multipurpose Love acoustic wave immunosensor for bacteria, virus or proteins detection. 1RBM29(2-3): 155-16]. Moll, N., Pascal, E., Dinh, D.H., Pillot, J.P., Bennetau, B., Rebiere, D. et al. (2007) A Love wave immunosensor for whole E. coli bacteria detection using an innovative two- step immobilisation approach. Biosensors & Bioelectronics 22(9-10): 2145-2150. Muhammad-Tahir, Z., and Alocilja, E.C. (2003a) A conductometric biosensor for biosecurity. Biosensors and Bioelectronics 18(5-6): 813-819. Muhammad-Tahir, Z., and Alocilja, E.C. (2003b) Fabrication of a disposable biosensor for Escherichia coli 0157:H7 detection. IEEE Sensors Journal 3(4): 345-351. Muhammad-Tahir, Z., Alocilja, EC, and Grooms, D.L. (2005) Rapid detection of Bovine viral diarrhea virus as surrogate of bioterrorism agents. IEEE Sensors Journal 5(4): 757-762. Nandakumar, V., La Belle, J .T., Reed, J., Shah, M., Cochran, D., Joshi, L., and Alford, TL. (2008) A methodology for rapid detection of Salmonella Typhimurium using label- free electrochemical impedance spectroscopy. Biosensors and Bioelectronics 24(4): 1039-1042. Nanduri, V., Kim, G., Morgarn, M.T., Ess, D., Hahm, B., Kothapalli, A. et a1. (2006) Antibody immobilization on waveguides using a flow-through system shows improved Listeria monocytogenes detection in an automated fiber optic biosensor: RAPTORTM. Sensors 6: 808-822. O'Sullivan, OK, and Guilbault, G.G. (1999) Commercial quartz crystal microbalances - theory and applications. Biosensors & Bioelectronics 14(8-9): 663-670. Ohira, M., Sakai, T., Takeuchi, M., Kobayashi, Y., and Tsuji, M. (1987) Raman and infrared spectra of polyaniline. Synthetic Metals 18(1-3): 347-352. Okinaka, R.T., Cloud, K., Hampton, O., Hoffmaster, A.R., Hill, K.K., Keim, P. et al. (1999) Sequence and organization of pXO], the large Bacillus anthracis plasmid harboring the anthrax toxin genes. Journal of Bacteriology 181(20): 6509-6515. 196 Pal, S., Alocilja, EC, and Downes, F.P. (2007) Nanowire labeled direct-charge transfer biosensor for detecting Bacillus species. Biosensors & Bioelectronics 22(9-10): 2329- 2336. Pal, S., Ying, W., Alocija, EC, and Downes, F.P. (2008) Sensitivity and specificity performance of a direct-charge transfer biosensor for detecting Bacillus cereus in selected food matrices. Biosystems Engineering 99(4): 461-468. Palchetti, I., and Mascini, M. (2008) Electroanalytical biosensors and their potential for food pathogen and toxin detection. Analytical and Bioanalytical Chemistry 391(2): 455- 47]. Pan, L.J., Pu, L., Shi, Y., Sun, T., Zhang, R., and Zheng, Y.D. (2006) Hydrothermal synthesis of polyaniline mesostructures. Advanced Functional Materials 16(10): 1279- 1288. Panda, B.R., Singh, A.K., Ramesh, A., and Chattopadhyay, A. (2008) Rapid Estimation of Bacteria by a Fluorescent Gold Nanoparticle-Polythiophene Composite. Langmuir 24(20): 11995-12000. Pannifer, A.D., Wong, T.Y., Schwarzenbacher, R., Renatus, M., Petosa, C., Bienkowska, J. et al. (2001) Crystal structure of the anthrax lethal factor. Nature 414(6860): 229-233. Park, M.E., and Chang, J.H. (2007) High throughput human DNA purification with aminosilanes tailored silica-coated magnetic nanoparticles. Materials Science & Engineering C-Biomimetic and Supramolecular Systems 27(5-8): 1232-1235. Patolsky, F., Zheng, G., and Lieber, CM. (2006) Nanowire-based biosensors. Analytical Chemistry 78(13): 4260-4269. Peruski, A.H., and Peruski, LP. (2003) Immunological methods for detection and identification of infectious disease and biological warfare agents. Clinical and Diagnostic Laboratory Immunology 10(4): 506-513. Petosa, C., Collier, R.J., Klimpel, K.R., Leppla, SH, and Liddington, RC. (1997) Crystal structure of the anthrax toxin protective antigen. Nature 385(6619): 833-838. 197 Pezard, C., Berche, P., and Mock, M. (1991) Contribution of Individual Toxin Components to Virulence of Bacillus anthracis. Infection and Immunity 59(10): 3472- 3477. Pileni,M.P. (2001) Semiconductor Nanocrystals. In Nanoscale Materials in Chemistry. Klabunde,K. (ed). John Wiley & Sons, Inc., pp. 85-120. Piletsky, S.A., Turner, N.W., and Laitenberger, P. (2006) Molecularly imprinted polymers in clinical diagnostics - Future potential and existing problems. Medical Engineering & Physics 28(10): 971-977. Pingarron, J .M., YBiez-Sedeio, P., and GonzBlez-Corth, A. (2008) Gold nanoparticle- based electrochemical biosensors. Electrochimica Acta 53(19): 5848-5 866. Poddar, P., Wilson, J.L., Srikanth, H., Morrison, SA, and Carpenter, BE. (2004b) Magnetic properties of conducting polymer doped with manganese-zinc ferrite nanoparticles. Nanotechnology 15(10): $570-$574. Poddar, P., Wilson, J.L., Srikanth, H., Morrison, SA, and Carpenter, BE. (2004a) Magnetic properties of conducting polymer doped with manganese-zinc ferrite nanoparticles. Nanotechnology 15(10): 8570-8574. Pouget, J .P., Jozefowicz, M.E., Epstein, A.J., Tang, X., and Macdiarmid, AG. (1991) X- Ray Structure of Polyaniline. Macromolecules 24(3): 779-789. Prabhakar, N., Arora, K., Singh, H., and Malhotra, B.D. (2008) Polyaniline based nucleic acid sensor. Journal of Physical Chemistry B 112(15): 4808-4816. Prabhakaran, T., and Hemalatha, J. (2008) Synthesis and characterization of magnetoelectric polymer nanocomposites. Journal of Polymer Science, Part B: Polymer Physics 46(22): 2418-2422. Prakash, R. (2002) Electrochemistry of polyaniline: Study of the pH effect and electrochromism. Journal of Applied Polymer Science 83(2): 378-3 85. Pumera, M., Castaneda, M.T., Pividori, M.l., Eritja, R., Merkoci, A., and Alegret, S. (2005) Magnetically Trigged Direct Electrochemical Detection of DNA Hybridization Using Au67 Quantum Dot as Electrical Tracer. Langmuir 21(21): 9625-9629. 198 Radke, S.M., and Alocilja, EC. (2005) A high density microelectrode array biosensor for detection of E. coli 0157:H7. Biosensors and Bioelectronics 20(8): 1662-1667. Rahman, M.A., Kumar, R, Park, D.S., and Shim, Y.B. (2008) Electrochemical sensors based on organic conjugated polymers. Sensors 8(1): 118-141. Ram, M.K., Salerno, M., Adami, M., Faraci, P., and Nicolini, C. (1999) Physical properties of polyaniline films: Assembled by the layer-by-layer technique. Langmuir 15(4): 1252-1259. Ramanathan, K., Bangar, M.A., Yun, M., Chen, W., Myung, N.V., and Mulchandani, A. (2005) Bioaffinity sensing using biologically functionalized conducting-polymer nanowire. Journal of the American Chemical Society 127(2): 496-497. Ramanavicius, A., Ramanaviciene, A., and Malinauskas, A. (2006) Electrochemical sensors based on conducting polymer- polypyrrole. Electrochimica Acta 51(27): 6025- 6037. Ray, A., Asturias, G.E., Kershner, D.L., Richter, A.F., Macdiarmid, AG, and Epstein, A.J. (1989) Polyaniline - Doping, Structure and Derivatives. Synthetic Metals 29(1): E141-E150. Reddy, K.R., Lee, K.P., Gopalan, AL, and Showkat, AM. (2007) Synthesis and properties of magnetite/poly (aniline-co-8-amino-2-naphthalenesulfonic acid) (SPAN) nanocomposites. Polymers for Advanced Technologies 18(1): 38-43. Reif, T.C., Johns, M., Pillai, S.D., and Carl, M. (1994) Identification of Capsule-Forming Bacillus anthracis Spores with the PCR and A Novel Dual-Probe Hybridization Format. Applied and Environmental Microbiology 60(5): 1622-1625. Rossi, A.M., Wang, L., Reipa, V., and Murphy, TE. (2007) Porous silicon biosensor for detection of viruses. Biosensors & Bioelectronics 23(5): 741-745. Rossi, L.M., Shi, L., Rosenzweig, N., and Rosenzweig, Z. (2006) Fluorescent silica nanospheres for digital counting bioassay of the breast cancer marker HER2/nue. Biosensors and Bioelectronics 21(10): 1900-1906. 199 Ruan, C.M., Zeng, K.F., Varghese, OK, and Grimes, CA. (2003) Magnetoelastic immunosensors: Amplified mass immunosorbent assay for detection of Escherichia coli 0157:H7. Analytical Chemistry 75(23): 6494-6498. Ryder, K.S., Morris, D.G., and Cooper, J.M. (1997) Role of conducting polymeric interfaces in promoting biological electron transfer. Biosensors & Bioelectronics 12(8): 721-727. Safina, G., van Lier, M., and Danielsson, B. (2008) Flow-injection assay of the pathogenic bacteria using lectin-based quartz crystal microbalance biosensor. Talanta 77(2): 468-472. Sai, V.V.R., Mahajan, S., Contractor, A.Q., and Mukherji, S. (2006) Immobilization of antibodies on polyaniline films and its application in a piezoelectric immunosensor. Analytical Chemistry 78(24): 8368-8373. Sandhu, A., Kumagai, Y., Lapicki, A., Sakamoto, S., Abe, M., and Handa, H. (2007) High efficiency Hall effect micro-biosensor platform for detection of magnetically labeled biomolecules. Biosensors and Bioelectronics 22(9-10): 2115-2120. Sapsford, K.E., Pons, T., Medintz, IL, and Mattoussi, H. (2006) Biosensing with luminescent semiconductor quantum dots. Sensors 6(8): 925-953. Sergeeva, T.A., Piletskii, S.A., Rachkov, AB, and El'skaya, A.V. (1996) Synthesis and Examination of Polyanilines as Labels in Immunosensor Analysis. Journal of Analytical Chemistry 51(4): 394. Shankaran, D.R., Gobi, K.V.A., and Miura, N. (2007) Recent advancements in surface plasmon resonance immunosensors for detection of small molecules of biomedical, food and environmental interest. Sensors and Actuators B-Chemical 121(1): 158-177. Sharma, R., Lamba, S., Annapoomi, S., Shanna, P., and Inoue, A. (2005) Composition dependent magnetic properties of iron oxide-polyaniline nanoclusters. Journal of Applied Physics 97(1). Shen, Z.H., Huang, M.C., Xiao, C.D., Zhang, Y., Zeng, X.Q., and Wang, PC. (2007) Nonlabeled quartz crystal microbalance biosensor for bacterial detection using carbohydrate and lectin recognitions. Analytical Chemistry 79(6): 2312-2319. 200 Singh, C., Agarwal, G.S., Rai, G.P., Singh, L., and Rao, V.K. (2005) Specific detection of Salmonella typhi using renewable amperometric immunosensor. Electroanalysis 17(22): 2062-2067. So, H.M., Park, D.W., Jeon, E.K., Kim, Y.H., Kim, HS, Lee, C.K. et al. (2008) Detection and titer estimation of Escherichia coli using aptamer-functionalized single- walled carbon-nanotube field-effect transistors. Small 4(2): 197-201. Song, L.N., Ahn, S., and Walt, DR. (2005) Detecting biological warfare agents. Emerging Infectious Diseases 1 1(10): 1629-1632. Song, L.N., Ahn, S., and Walt, DR. (2006) Fiber-optic microsphere-based arrays for multiplexed biological warfare agent detection. Analytical Chemistry 78(4): 1023-1033. Song, S.P., Wang, L.H., Li, J., Zhao, J.L., and Fan, CH. (2008) Aptamer-based biosensors. Trac-Trends in Analytical Chemistry 27(2): 108-1 17. Sorensen, CM. (2001) Magnetism. In Nanoscale Materials in Chemistry. Klabunde,K. (ed). John Wiley & Sons, Inc., pp. 169-222. Sotiropoulou, S., Vamvakaki, V., and Chaniotakis, NA. (2005) Stabilization of enzymes in nanoporous materials for biosensor applications. Biosensors & Bioelectronics 20(8): 1674-1679. Soukka, T., Harma, H., Paukkunen, J ., and Lovgren, T. (2001) Utilization of Kinetically Enhanced Monovalent Binding Affinity by Immunoassays Based on Multivalent Nanoparticle-Antibody Bioconjugates. Analytical Chemistry 73(10): 2254-2260. Spanier, J .E. (2006) One-dimensional semiconductor and oxide nanostructures. In Nanomaterials Handbook. Gogotsi,Y. (ed). Taylor and Francis, pp. 283-316. Spencer, RC. (2003) Bacillus anthracis. Journal of Clinical Pathology 56(3): 182-187. Stafstrom, S., Bredas, J.L., Epstein, A.J., Woo, H.S., Tanner, D.B., Huang, W.S., and Macdiarmid, AC. (1987) Polaron Lattice in Highly Conducting Polyaniline - Theoretical and Optical Studies. Physical Review Letters 59(13): 1464-1467. 201 Stejskal, J ., and Gilbert, RC. (2002) Polyaniline. Preparation of a conducting polymer (IUPAC technical report). Pure and Applied Chemistry 74(5): 857-867. Stej skal, J ., Kratochvil, P., and Jenkins, AD. (1996) The formation of polyaniline and the nature of its structures. Polymer 37(2): 367-369. Stringer, R.C., Schommer, S., Hoehn, D., and Grant, SA. (2008) Development of an optical biosensor using gold nanoparticles and quantum dots for the detection of Porcine Reproductive and Respiratory Syndrome Virus. Sensors and Actuators B-Chemical 134(2): 427-431. Su, X.L., and Li, Y.B. (2004a) A self-assembled monolayer-based piezoelectric immunosensor for rapid detection of Escherichia coli 0157:H7. Biosensors & Bioelectronics 19(6): 563-574. Su, X.L., and Li, Y.B. (2004b) Quantum dot biolabeling coupled with immunomagnetic separation for detection of Escherichia coli 0157:H7. Analytical Chemistry 76(16): 4806-4810. Subramanian, A., Irudayaraj, J ., and Ryan, T. (2006) A mixed self-assembled monolayer- based surface plasmon immunosensor for detection of E. coli 0157:H7. Biosensors & Bioelectronics 21 (7): 998-1006. Tahir, Z.M., Alocilja, EC, and Grooms, D.L. (2007) Indium tin oxide-polyaniline biosensor: Fabrication and characterization. Sensors 7(7): 1123-1140. Taitt, C.R., Anderson, G.P., Lingerfelt, B.M., Feldstein, M.J., and Ligler, ES. (2002) Nine-analyte detection using an array-based biosensor. Analytical Chemistry 74(23): 6114-6120. Tamanaha, C.R., Mulvaney, S.P., Rife, J.C., and Whitman, L.J. (2008) Magnetic labeling, detection, and system integration. Biosensors and Bioelectronics 24(1): 1-13. Tang, D., Yuan, R., and Chai, Y. (2008) Ultrasensitive Electrochemical Immunosensor for Clinical Immunoassay Using Thionine-Doped Magnetic Gold Nanospheres as Labels and Horseradish Peroxidase as Enhancer. Analytical Chemistry 80(5): 1582-1588. 202 Tang, H., Kitani, A., and Shiotani, M. (1996) Cyclic voltammetry of K1 at polyaniline- filmed Pt electrodes Part 1: formation of polyaniline-iodine charge transfer complexes. Journal of Applied Electrochemistry 26(1): 36-44. Taylor, A.D., Ladd, J ., Yu, Q., Chen, S., Homola, J., and Jiang, S. (2006) Quantitative and simultaneous detection of four foodbome bacterial pathogens with a multi-channel SPR sensor. Biosensors and Bioelectronics 22(5): 752-758. Terry, L.A., White, S.F., and Tigwell, L.J. (2005) The application of biosensors to fresh produce and the wider food industry. Journal of A gricultural and Food Chemistry 53(5): 1309-1316. Theegala, C.S., Small, DD, and Monroe, W.T. (2008) Oxygen electrode-based single antibody amperometric biosensor for qualitative detection of E. coli and bacteria in water. Journal of Environmental Science and Health Part A-Toxic/Hazardous Substances & Environmental Engineering 43(5): 478-487. Thurer, R., Vigassy, T., Hirayama, M., Wang, J., Bakker, E., and Pretsch, E. (2007) Potentiometric immunoassay with quantum dot labels. Analytical Chemistry 79(13): 5107-5110. Tims, TB, and Lim, D.V. (2004) Rapid detection of Bacillus anthracis spores directly from powders with an evanescent wave fiber-optic biosensor. Journal of Microbiological Methods 59(1): 127-130. Vaisocherova, H., Mrkvova, K., Piliarik, M., Jinoch, P., Steinbachova, M., and Homola, J. (2007) Surface plasmon resonance biosensor for direct detection of antibody against Epstein-Baff virus. Biosensors & Bioelectronics 22(6): 1020-1026. Varshney, M., and Li, Y. (2007) Interdigitated array microelectrode based impedance biosensor coupled with magnetic nanoparticle-antibody conjugates for detection of Escherichia coli 0157:H7 in food samples. Biosensors and Bioelectronics 22(11): 2408- 2414. Varshney, M., Yang, L., Su, X.L., and Li, Y. (2005) Magnetic Nanoparticle-Antibody Conjugates for the Separation of Escherichia coli 0157:H7 in Ground Beef. Journal of Food Protection 68: 1804-1811. Vaseashta, A., and mova-Malinovska, D. (2005) Nanostructured and nanoscale devices, sensors and detectors. Science and Technology of Advanced Materials 6(3-4): 312-318. 203 Villamizar, R.A., Maroto, A., Rius, F.X., Inza, I., and Figueras, M.J. (2008) Fast detection of Salmonella Infantis with carbon nanotube field effect transistors. Biosensors & Bioelectronics 24(2): 279-283. Waggoner, PS, and Craighead, H.G. (2007) Micro- and nanomechanical sensors for environmental, chemical, and biological detection. Lab on A Chip 7(10): 1238-1255. Wan, M., and Li, J. (1998) Synthesis and electrical-magnetic properties of polyaniline composites. Journal of Polymer Science, Part A: Polymer Chemistry 36(15): 2799-2805. Wang, J. (2005) Nanomaterial-based electrochemical biosensors. Analyst 130(4): 421- 426. Wang, J ., Liu, G., and Lin, Y. (2007a) Nanotubes, Nanowires, and Nanocantilevers in Biosensor Development. In Nanomaterials for Biosensors. Kumar,C.S.S.R. (ed). Wiley- VCH, pp. 56-100. Wang, J. (2002) Electrochemical nucleic acid biosensors. Analytica Chimica Acta 469(1): 63-71. Wang, L., Zhao, W.J., O'Donoghue, M.B., and Tan, W.H. (2007b) Fluorescent nanoparticles for multiplexed bacteria monitoring. Bioconjugate Chemistry 18(2): 297- 30]. Wang, L.J., Wei, Q.S., Wu, C.S., Hu, Z.Y., Ji, J., and Wang, P. (2008) The Escherichia coli 0157:H7 DNA detection on a gold nanoparticle-enhanced piezoelectric biosensor. Chinese Science Bulletin 53(8): 1175-1184. Wang, L., Wei, Q., Wu, C., Ji, J., and Wang, P. A QCM biosensor based on gold nanoparticles amplification for real-time bacteria DNA detection. International Conference on Information Acquisition, ICIA 2007, July 09,2007 - July 11,2007. 46-51. 2007c. Jeju City, Korea, Republic of, Inst. of Elec. and Elec. Eng. Computer Society. Proceedings of the 2007 International Conference on Information Acquisition, ICIA. Wang, S.X., Bae, S.Y., Li, G.X., Sun, S.H., White, R.L., Kemp, J.T., and Webb, CD. (2005) Towards a magnetic microarray for sensitive diagnostics. Journal of Magnetism and Magnetic Materials 293(1): 731-736. 204 Wang, S.X., and Li, G. (2008) Advances in giant magnetoresistance biosensors with magnetic nanoparticle tags: Review and outlook. IEEE Transactions on Magnetics 44(7): 1687-1702. Waswa, J ., Irudayaraj, J., and DebRoy, C. (2007) Direct detection of E coli 0157:H7 in selected food systems by a surface plasmon resonance biosensor. Lwt-Food Science and Technology 40(2): 187-192. Waswa, J.W., DebRoy, C., and Irudayaraj, J. (2006) Rapid detection of Salmonella Enteritidis and Escherichia coli using surface plasmon resonance biosensor. Journal of Food Process Engineering 29(4): 373-385. Wei, D., and Ivaska, A. (2006) Electrochemical biosensors based on polyaniline. Chemia Analityczna 51(6): 839-852. Wei, D., Baral, J .K., Osterbacka, R., and Ivaska, A. (2008) Electrochemical fabrication of a nonvolatile memory device based on polyaniline and gold particles. Journal of Materials Chemistry 18(16): 1853-1857. Wu, T.Z., Su, C.C., Chen, L.K., Yang, H.H., Tai, DR, and Peng, KC. (2005) Piezoelectric immunochip for the detection of dengue fever in viremia phase. Biosensors & Bioelectronics 21(5): 689-695. Xu, C.J., and Sun, SH. (2007) Monodisperse magnetic nanoparticles for biomedical applications. Polymer International 56(7): 821-826. Xue, W.Y., Qiu, H., Fang, K., Li, J., Zhao, J.W., and Li, M. (2006) Electrical and magnetic properties of the composite pellets containing DBSA-doped polyaniline and Fe nanoparticles. Synthetic Metals 156(11-13): 833-837. Yang, L.J., and Li, Y.B. (2006) Detection of viable Salmonella using microelectrode- based capacitance measurement coupled with immunomagnetic separation. Journal of Microbiological Methods 64(1): 9-16. Yao, C.Y., Zhu, T.Y., Tang, J ., Wu, R., Chen, Q.H., Chen, M. et al. (2008) Hybridization assay of hepatitis B virus by QCM peptide nucleic acid biosensor. Biosensors & Bioelectronics 23(6): 879-885. 205 Ybarra, G., Moina, C., Florit, M.l., and Posadas, D. (2000) Proton exchange during the redox switching of polyaniline film electrodes. Electrochemical and Solid-State Letters 3(7): 330-332. Ye, L., and Haupt, K. (2004) Molecularly imprinted polymers as antibody and receptor mimics for assays, sensors and drug discovery. Analytical and Bioanalytical Chemistry 378(8): 1887-1897. Yeung, S.W., and Hsing, I.M. (2006) Manipulation and extraction of genomic DNA from cell lysate by functionalized magnetic particles for lab on a chip applications. Biosensors & Bioelectronics 21(7): 989-997. Yoon, H., Kim, J.H., Lee, N., Kim, B.G., and Jang, J. (2008) A novel sensor platform based on aptamer-conjugated polypyrrole nanotubes for label-free electrochemical protein detection. Chembiochem 9(4): 634-641. Yu, C.X., and Irudayaraj, J. (2007) Multiplex biosensor using gold nanorods. Analytical Chemistry 79(2): 572-579. Yun, Y., Bange, A., Heineman, W.R., Halsall, H.B., Shanov, V.N., Dong, Z. et al. (2007) A nanotube array immunosensor for direct electrochemical detection of antigen-antibody binding. Sensors and Actuators B: Chemical 123(1): 177-182. Zhang, C.Y., Yeh, H.C., Kuroki, M.T., and Wang, T.H. (2005a) Single-quantum-dot- based DNA nanosensor. Nature Materials 4(11): 826-831. Zhang, D., Carr, D.J., and Alocilja, E.C. (2009a) Fluorescent bio-barcode DNA assay for the detection of Salmonella enterica serovar Enteritidis. Biosensors and Bioelectronics 24(5): 1377-1381. Zhang, L., Zhang, J ., and Zhang, C. (2009b) Electrochemical synthesis of polyaniline nano-network on α-alanine functionalized glassy carbon electrode and its application for the direct electrochemistry of horse heart cytochrome c. Biosensors and Bioelectronics 24(7): 2085-2090. Zhang, Z.M., Wan, M.X., and Wei, Y. (2005b) Electromagnetic fimctionalized polyaniline nanostructures. Nanotechnology 16(12): 2827-2832. 206 Zhang, Z., and Wan, M. (2003) Nanostructures of polyaniline composites containing nano-magnet. Synthetic Metals 132(2): 205-212. Zhao, G., Xing, F., and Deng, S. (2007) A disposable amperometric enzyme immunosensor for rapid detection of Vibrio parahaemolyticus in food based on agarose/Nano-Au membrane and screen-printed electrode. Electrochemistry Communications 9(6): 1263-1268. Zhao, W., Brook, M.A., and Li, Y.F. (2008) Design of Gold Nanoparticle-Based Colorimetric Biosensing Assays. Chembiochem 9(15): 2363-2371. Zhao, X.J., Hilliard, L.R., Mechery, S.J., Wang, Y.P., Bagwe, R.P., Jin, 8.6., and Tan, W.H. (2004) A rapid bioassay for single bacterial cell quantitation using bioconjugated nanoparticles. Proceedings of the National Academy of Sciences of the United States of America 101(42): 15027-15032. Zhou, J., Tsao, H.K., Sheng, Y.J., and Jiang, S.Y. (2004) Monte Carlo simulations of antibody adsorption and orientation on charged surfaces. Journal of Chemical Physics 121(2): 1050-1057. Zhou, R.H., Wang, P., and Chang, HQ (2006) Bacteria capture, concentration and detection by alternating current dielectrophoresis and self-assembly of dispersed single- wall carbon nanotubes. Electrophoresis 27(7): 1376-1385. Zhu, L., Yang, R., Zhai, J ., and Tian, C. (2007) Bienzymatic glucose biosensor based on co-immobilization of peroxidase and glucose oxidase on a carbon nanotubes electrode. Biosensors and Bioelectronics 23(4): 528-535. Zhu, N.N., Chang, Z., He, P.G., and Fang, Y.Z. (2006a) Electrochemically fabricated polyaniline nanowire-modified electrode for voltammetric detection of DNA hybridization. Electrochimica Acta 51(18): 375 8-3 762. Zhu, N.N., Zhang, A.P., He, P.G., and Fang, Y.Z. (2004) DNA hybridization at magnetic nanoparticles with electrochemical stripping detection. Electroanalysis 16(23): 1925- 1930. Zhu, X., Han, K., and Li, G. (2006b) Magnetic Nanoparticles Applied in Electrochemical Detection of Controllable DNA Hybridization. Analytical Chemistry 78(7): 2447-2449. 207 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII 1111111111111111111111111111111ll