Low-Cost, Rapid, Sensitive Detection of Pathogenic Bacteria Using Phage-Based Magnetoelastic Biosensors by Shin Horikawa A dissertation submitted to the Graduate Faculty of Auburn University in partial ful llment of the requirements for the Degree of Doctor of Philosophy Auburn, Alabama May 5, 2013 Keywords: magnetoelastic biosensor, phage, pathogenic bacteria, food safety, biosecurity Copyright 2013 by Shin Horikawa Approved by Bryan A. Chin, Chair, Professor of Materials Engineering Zhongyang Cheng, Professor of Materials Engineering Dong-Joo Kim, Associate Professor of Materials Engineering Valery A. Petrenko, Professor of Pathobiology Sang-Jin Suh, Associate Professor of Biological Sciences George T. Flowers, Dean of the Graduate School Abstract As part of the ongoing e orts to secure food safety as well as to guard against possible bioterrorism, the role of pathogen detection technologies has become vi- tal. However, conventional and standard detection methods, including culture-, immunology-, and polymerase chain reaction-based methods, are generally expen- sive, time-consuming, and labor-intensive. Hence, there is a need for new detection technologies that outperform the conventional methods and enable the rapid, on-site detection of pathogenic substances. Although label-free biosensors have proven to be among the most promising methods, meeting various performance criteria (e.g., sen- sitivity, selectivity, assay time, thermal stability, and longevity) simultaneously still remains a challenge. Hence, further research and development are essential before biosensors become a reliable, alternative solution. Phage-based magnetoelastic (ME) biosensors, a novel class of wireless, mass- sensitive biosensors, are among potential candidates that could overcome the above performance challenge. These biosensors are not only thermally robust, but their wireless nature of detection o ers great exibility in design and use, which facili- tates on-site pathogen detection. In addition, the sensitivity of ME biosensors can be improved by reducing their dimensions, and the fabrication cost per sensor can be reduced via batch fabrication. Hence, this dissertation presents investigations into the performance improvement of phage-based ME biosensors, in terms of cost- e ectiveness, rapidness, and sensitivity, and into the enhanced detection of pathogenic bacteria, Salmonella Typhimurium and Bacillus anthracis spores, for food safety and biosecurity. ii To enhance both cost-e ectiveness and sensitivity, micron- to millimeter-scale ME biosensors were batch-fabricated and used. In this way, the fabrication cost per sensor was reduced to a fraction of a cent. In addition, the following two method- ologies were employed to dramatically shorten assay time: (1) direct detection of S. Typhimurium on fresh spinach leaves and (2) detection of B. anthracis spores with the aid of a designed micro uidic ow cell, which ensures e cient physical contact between a biosensor and owing spores. By using these methodologies with low-cost, miniature ME biosensors, (1) S. Typhimurium cells on the order of 104 cells/cm2 were detected with 150- m long sensors in 45 min, and (2) down to 106 B. anthracis spores were detected with 200- m long sensors in 10 min. Additionally, to further enhance the detection capabilities of phage-based ME biosensors, the following e ects were studied: (1) the e ects of mass position on the sensitivity of ME biosensors and (2) the e ects of surface functionalization on surface phage coverage. The mass sensitivity of ME biosensors was found to be largely dependent on the dimensions of the sensors as well as on the position of attached masses. From numerical simulation results, a formula that predicts the mass-position-dependent sensor response for a single localized mass was also derived. In addition, surface phage coverage on bare and surface-functionalized ME biosen- sors was quanti ed by atomic force microscopy. The results showed that activated carboxyl-based covalent attachment produced a surface phage coverage of 50%, which is comparable to that obtained through physical adsorption, the traditional method of phage immobilization. By contrast, much lower surface phage coverages ( 5%) were obtained for aldehyde- and methyl-terminated sensor surfaces. These di erences in surface phage coverage was also found to a ect the quantity of a sub- sequently captured analyte. Hence, by properly functionalizing the sensor surface, both surface phage coverage and the quantity of the captured analyte can be con- trolled. Finally, with the results of the mass-position-dependence of sensor response, iii a concept of phage layer patterning was introduced. Phage may be patterned onto desired parts of the sensor surface to further enhance the detection capabilities of ME biosensors. iv Acknowledgments This dissertation would not have been possible without the guidance and help of a great number of individuals. First and foremost, I am truly indebted to my advisor, Dr. Bryan A. Chin, who has provided invaluable assistance in the preparation and completion of this research. I will never forget his generosity and encouragement. I am also grateful to all my committee members, Dr. Zhongyang Cheng, Dr. Dong-Joo Kim, Dr. Valery A. Petrenko, and Dr. Sang-Jin Suh for their sincere support. Special thanks to Dr. Valery A. Petrenko, who has helped me improve my attitude towards scienti c research. I would like to thank Dr. Maria L. Auad, Dr. Michael J. Bozack, and Dr. Michael E. Miller for their technical assistance in atomic force microscopy, x-ray photoelectron spectroscopy, and confocal scanning laser microscopy, respectively. With their expert guidance and assistance, I have been able to acquire sound engineering practices. I am also thankful for all the support from Dr. James M. Barbaree, Kiril A. Vaglenov, and I-Hsuan Chen, who have provided all biological samples and shared a great deal of knowledge regarding microbiology over the past several years. Yating Chai, Dr. John Shu, Michael L. Johnson, Leslie C. Mathison, Dr. Shichu Huang, Dr. Jiehui Wan, Dr. Suiqiong Li, and Steve Best have shared with me countless discussions on sensing principles, microelectronic fabrication, and testing methodologies. We have worked together in our experiments and constantly tried to gain greater understanding of our work. I would like to thank them for always being there for me. Last but not least, I owe earnest thankfulness to my parents and sister for their love and support. v Table of Contents Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xix List of Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxi 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1 Background and need . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Research objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 1.3 Target pathogenic bacteria to be detected . . . . . . . . . . . . . . . 6 1.3.1 Salmonella Typhimurium . . . . . . . . . . . . . . . . . . . . 6 1.3.2 Bacillus anthracis spores . . . . . . . . . . . . . . . . . . . . . 8 1.4 Organization of this dissertation . . . . . . . . . . . . . . . . . . . . . 10 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2 Review of the Literature on Bacterial Detection Methods . . . . . . . . . 14 2.1 Conventional detection methods . . . . . . . . . . . . . . . . . . . . . 14 2.1.1 Culture-based methods . . . . . . . . . . . . . . . . . . . . . . 14 2.1.2 Immunology-based methods . . . . . . . . . . . . . . . . . . . 15 2.1.3 Polymerase chain reaction-based methods . . . . . . . . . . . 17 2.2 Biosensors as promising bacterial detection methods . . . . . . . . . . 20 2.2.1 De nition of a biosensor . . . . . . . . . . . . . . . . . . . . . 20 2.2.2 Biomolecular-recognition elements . . . . . . . . . . . . . . . . 22 2.2.3 Signal transducers . . . . . . . . . . . . . . . . . . . . . . . . 23 2.3 Conventional detection methods vs. biosensors . . . . . . . . . . . . . 23 vi 2.3.1 Probability of detection: PCR vs. biosensors . . . . . . . . . . 25 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 3 Phage-Based Magnetoelastic (ME) Biosensors . . . . . . . . . . . . . . . 37 3.1 Landscape phages as biomolecular-recognition elements . . . . . . . . 37 3.2 Magnetoelasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 3.2.1 Joule magnetostriction . . . . . . . . . . . . . . . . . . . . . . 42 3.2.2 Magnetization and Joule magnetostriction in ferromagnets . . 44 3.2.3 ME signal transducers . . . . . . . . . . . . . . . . . . . . . . 47 3.3 Fabrication of ME sensor platforms . . . . . . . . . . . . . . . . . . . 48 3.3.1 Dicing method . . . . . . . . . . . . . . . . . . . . . . . . . . 48 3.3.2 Co-sputtering-based method . . . . . . . . . . . . . . . . . . . 49 3.3.3 Annealing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 3.3.4 Fabrication cost per sensor platform . . . . . . . . . . . . . . . 53 3.4 Fabrication of phage-based ME biosensors . . . . . . . . . . . . . . . 55 3.4.1 Immobilization of a phage on the ME sensor platforms . . . . 55 3.4.2 Surface blocking of the ME biosensors with bovine serum albumin 55 3.5 Principle of detection . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 3.5.1 Minimum detectable number of bacterial cells . . . . . . . . . 57 3.5.2 Measurement of the resonant frequency of the ME biosensors . 59 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 4 Direct Detection of S. Typhimurium on Fresh Spinach Leaves . . . . . . 64 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 4.2 Material and methods . . . . . . . . . . . . . . . . . . . . . . . . . . 66 4.2.1 E2 phage and S. Typhimurium . . . . . . . . . . . . . . . . . 66 4.2.2 Confocal re ectance imaging of spinach leaf surfaces . . . . . . 66 4.2.3 Fabrication of ME sensor platforms with three di erent sizes . 67 4.2.4 Fabrication of phage-based ME biosensors . . . . . . . . . . . 68 vii 4.2.5 Determination of the concentration of BSA for surface blocking 68 4.2.6 Direct detection of S. Typhimurium on fresh spinach leaves . . 69 4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 4.3.1 Observation of Salmonella-inoculated leaf surfaces . . . . . . . 71 4.3.2 Resonant frequency measurement . . . . . . . . . . . . . . . . 72 4.3.3 Dose-response of the ME biosensors . . . . . . . . . . . . . . . 74 4.3.4 Determination of the LOD . . . . . . . . . . . . . . . . . . . . 76 4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 4.4.1 Topography of leaf surfaces and its e ects on the LOD . . . . 79 4.4.2 E ects of the number of biosensors on the LOD . . . . . . . . 83 4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 5 Detection of B. anthracis Spores with the Aid of A Micro uidic Flow Cell 95 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 5.2 Material and methods . . . . . . . . . . . . . . . . . . . . . . . . . . 97 5.2.1 JRB7 phage and B. anthracis Sterne spores . . . . . . . . . . 97 5.2.2 Batch-fabrication of micron-scale ME resonators . . . . . . . . 97 5.2.3 Micro uidic ow cells . . . . . . . . . . . . . . . . . . . . . . . 98 5.2.3.1 Design and fabrication of the Type I micro uidic ow cell . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 5.2.3.2 Design and fabrication of the Type II micro uidic ow cell . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 5.3 Valve actuation and sample injection . . . . . . . . . . . . . . . . . . 104 5.4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 105 5.4.1 Con rmation of valve actuation . . . . . . . . . . . . . . . . . 105 5.4.2 Manipulation of uorescent-labeled micro-spheres . . . . . . . 105 5.4.3 Manipulation of a sensor . . . . . . . . . . . . . . . . . . . . . 106 viii 5.4.4 Detection of B. anthracis spores with the Type II micro uidic ow cell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 6 Enhancing the Detection Capabilities of Phage-Based ME Biosensors . . 115 6.1 E ects of mass position on the mass sensitivity of ME biosensors . . . 115 6.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 6.1.2 Material and methods . . . . . . . . . . . . . . . . . . . . . . 116 6.1.2.1 Microcontact printing . . . . . . . . . . . . . . . . . 116 6.1.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . 117 6.1.3.1 Finite element modal simulation . . . . . . . . . . . 118 6.2 E ects of surface functionalization on surface phage coverage . . . . . 123 6.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 6.2.2 Material and methods . . . . . . . . . . . . . . . . . . . . . . 125 6.2.2.1 Preparation of biological samples . . . . . . . . . . . 126 6.2.2.2 Manufacture of gold-coated ME resonators . . . . . . 126 6.2.2.3 Surface functionalization of gold-coated ME resonators 126 6.2.2.4 X-ray photoelectron spectroscopy . . . . . . . . . . . 127 6.2.2.5 Loading of the phage on the ME resonators . . . . . 128 6.2.2.6 Resonant frequency measurement . . . . . . . . . . . 128 6.2.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . 129 6.2.3.1 Veri cation of the surface functionalization of gold . 129 6.2.3.2 Surface phage coverage . . . . . . . . . . . . . . . . . 131 6.2.3.3 SEM observation and dose-response results . . . . . . 135 6.2.3.4 Enhancing the detection capabilities of ME biosensors with a patterned phage layer . . . . . . . . . . . . . 138 6.2.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139 ix Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140 7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142 x List of Figures 1.1 Economic cost - bene t assessment. . . . . . . . . . . . . . . . . . . . . . 2 1.2 Civilian biodefense funding by scal year, FY2001 - FY2013 (in a36millions). 3 1.3 S. Typhimurium cells on a spinach leaf surface. . . . . . . . . . . . . . . 8 1.4 B. anthracis spores on a gold surface. . . . . . . . . . . . . . . . . . . . . 9 2.1 Conventional methods for bacterial detection. . . . . . . . . . . . . . . . 14 2.2 Typical procedure for sandwich ELISA. . . . . . . . . . . . . . . . . . . 16 2.3 Schematic illustration of the PCR cycle. . . . . . . . . . . . . . . . . . . 18 2.4 Number of research articles published between 1985 and 2005 on di erent bacterial detection methods. . . . . . . . . . . . . . . . . . . . . . . . . . 20 2.5 Schematic diagram of a biosensor. . . . . . . . . . . . . . . . . . . . . . . 21 2.6 Classi cation of biosensors. . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.7 Probability of detection with respect to the inhalation dose of the target pathogen for PCR-based detectors. The following values were used in the calculations: Ws = 1,000 l/min, Ke = 0.8, I = 15, and m = 1. . . . . . . 28 2.8 Probability of detection with respect to the inhalation dose of the target pathogen for biosensor-based detectors. The following values were used in the calculations: Ws = 1,000 l/min, Ke = 0.8, Vs = 1 ml, and n = 1. . . 29 xi 3.1 Schematic illustration of the wild-type fd phage and its genetically engi- neered form, displaying a foreign peptide on the major coat protein pVIII. 38 3.2 Sequences of amino acid residues of the fd coat proteins. The N-terminus is to the left. The hydrophobic domains are underlined, whereas charged residues are indicated by + or -. . . . . . . . . . . . . . . . . . . . . . . . 39 3.3 Sequences of amino acid residues of the wild-type and fusion pVIII proteins. 41 3.4 Joule magnetostriction of a spherical ME material. . . . . . . . . . . . . 43 3.5 Hypothetical eld dependencies of k and ?. . . . . . . . . . . . . . . . 44 3.6 Magnetic domains and magnetization processes in a ferromagnet. . . . . 45 3.7 (a) E ects of magnetizing eld and mechanical stress on the distribution of magnetic moments in a ferromagnet and (b) eld dependence of k under a compressive stress. . . . . . . . . . . . . . . . . . . . . . . . . . 46 3.8 Temperature dependence of normalized s in Fe80B20. . . . . . . . . . . . 46 3.9 Diced sensor platforms stored in dry methanol. . . . . . . . . . . . . . . 49 3.10 Procedure for the co-sputtering-based method. . . . . . . . . . . . . . . . 50 3.11 Denton sputter coater. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 3.12 Scanning electron micrograph of batch-fabricated sensor platforms with a size of 100 m 25 m 4 m on a gold-coated wafer. . . . . . . . . . 52 3.13 Uniform attachment of bacterial cells or spores on a phage-immobilized ME biosensor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 3.14 Measurement setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 xii 3.15 Response of a typical 500 m 100 m 4 m ME biosensor in air. . . 60 4.1 Scanning electron micrographs of various produce surfaces: (a) tomato, (b) eggshell, and (c) spinach leaf. A close-up view of a spinach leaf spiked with S. Typhimurium is shown in (d). . . . . . . . . . . . . . . . . . . . 65 4.2 Di erently sized sensor platforms used (top view). . . . . . . . . . . . . . 67 4.3 E ects of BSA concentration on (a) resonant frequency changes for mea- surement and control sensors (2-mm long) and on (b) the con dence level of di erence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 4.4 Schematic illustration of the test procedure: (a) spot-inoculation of S. Typhimurium on the leaf surface and measurement of the initial resonant frequency of biosensors, (b) placement of both measurement and control sensors on the Salmonella-inoculated sites (after drying the Salmonella drops and misting the leaf surface), (c) measurement of the nal resonant frequency of the biosensors, and (d) typical responses of the biosensors. . 70 4.5 Scanning electron micrographs of a spinach leaf surface inoculated with a 40- l drop of S. Typhimurium with various concentrations: (a) 5 108 cells/ml, (b) 5 107 cells/ml, (c) 5 106 cells/ml (with a 150 m-long ME biosensor), and (d) 0 cells/ml (reference). . . . . . . . . . . . . . . . 71 4.6 Response of a typical 150 m 30 m 4 m sensor in air: (a) raw data set (10-time averaged) and its smoothed curve and (b) Lorentizan tting of the smoothed curve. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 4.7 Resonant peaks for a 150- m long sensor before and after placing on a leaf surface inoculated with S. Typhimurium at a concentration of 5 108 cells/ml. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 xiii 4.8 Dose-response plots for the di erently sized biosensors (2 mm-, 0.5 mm- and, 150 m-long sensors). The plots on the left and right are the results for the adaxial and abaxial surfaces, respectively. . . . . . . . . . . . . . 75 4.9 (a) Sigmoidal curve and (b) the determination of the LOD. . . . . . . . . 76 4.10 Background-subtracted data for the dose-response plots in Fig. 4.8. These data were tted with sigmoidal functions (red solid curves). The R2 values were all close to one. The values of fAVE + 3 are shown in blue text. 78 4.11 Typical height maps (a & b) and associated averaged pro les (c & d) of a leaf surface obtained along di erent sampling lengths. A three- dimensional representation of a leaf surface expressed by Eq. 4.2 is shown in (e). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 4.12 Finite well model of a Salmonella-inoculated leaf surface. The total sur- face area is x2LW: There are three types of wells: sensor-containing wells, cell-containing wells, and empty wells. . . . . . . . . . . . . . . . . . . . 84 4.13 Finite well model with n = 1 (i.e., one cell-containing well). . . . . . . . 85 4.14 Probability of detection with respect to the number of biosensors, m, and the number of cell-containing wells, n. Biosensors with lateral dimensions of 50 m 10 m were placed on a leaf surface of 0.26 cm2. . . . . . . . 87 4.15 Dependence of the LOD on the number of biosensors (50 m 10 m 2 m), m, for various values of P(D). . . . . . . . . . . . . . . . . . . . . 90 5.1 Scanning electron micrographs of 200 m 40 m 4 m ME resonators fabricated on a at wafer: (a) batch-fabricated resonators and (b) a close- up view. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 xiv 5.2 Procedure for the fabrication of a micro udic chip by multilayer soft lithography. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 5.3 Design of the Type I ow cell: (a) a three-dimensional view of the whole chip and (b) a close-up view of the key elements of the chip. . . . . . . . 100 5.4 Push-up and push-down valves. . . . . . . . . . . . . . . . . . . . . . . . 102 5.5 Design of the Type II ow cell: (a) a three-dimensional view of the whole chip on a slotted glass slide and (b) a close-up view of the key elements of the chip. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 5.6 A close-up top view of a fabricated chip (Type I), connected to external pressure sources for valve actuation and sample injection. . . . . . . . . . 104 5.7 Close-up top views of a Type I ow cell: (a) All the valves (#1 through #4) are closed, and (b) only the #2 and #4 valves are closed such that the green dye injected through the vertical channels can only ll out the center chamber. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 5.8 Fluorescent micrographs showing that the ow cell can separate a few micro-spheres into the reaction chamber: (a) separation of four spheres, (b) two spheres, and (c) one sphere. . . . . . . . . . . . . . . . . . . . . 106 5.9 Injection of a 200 m 40 m 4 m sensor into the chamber through the horizontal channels. . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 5.10 Optical micrograph of the Type II micro uidic ow cell. All the push- down valves are closed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 5.11 Fluorescent micrograph showing B. anthracis spores are bound on a 200 m 40 m 4 m ME biosensor. The chamber was outlined with a yellow solid line for better visualization. . . . . . . . . . . . . . . . . . . 109 xv 5.12 Streamlines in the micro uidic ow cell. The color bar below indicates the magnitude of ow velocity. . . . . . . . . . . . . . . . . . . . . . . . 110 5.13 Responses of ME biosensors (200 m 40 m 4 m) to various numbers of spores. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 6.1 PDMS stamps and microcontact printing. SA beads (not to scale) were placed in the middle (stamping area: 1 mm 0.8 mm) or (b) at both ends (stamping area: 0.5 mm 0.8 mm each) of a phage-immobilized ME biosensor using the Type A or Type B stamp, respectively. BSA is not shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 6.2 Phage-based ME biosensors loaded with SA beads. Beads are attached (a) in the middle, (b) at both ends, or (c) uniformly on both sides of the ME biosensors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 6.3 Three-dimensional FE model: (a) geometry, (b) meshed geometry, and (c) resultant mode shape. . . . . . . . . . . . . . . . . . . . . . . . . . . 119 6.4 Dependence of the mass sensitivity, f= m, on the longitudinal position of the attached mass and on the dimensions of the ME biosensors. . . . . 121 6.5 Surface functionalization of gold (1 1 1) with three SAM chemicals, based on the sulfur - gold chemistry: (a) AC (activated carboxyl-terminated), (b) ALD (aldehyde-terminated), and (c) MT (methyl-terminated). . . . . 125 xvi 6.6 Schematic illustration of the frequency measurement setup for ME biosen- sors. (a) The setup consists of a copper solenoid coil, a bar magnet, and a network analyzer (not shown). (b) A phage-immobilized ME biosensor is placed in the glass capillary ow cell and positioned in the coil center. A suspension of SA beads was passed at 25 l/min over the sensor, and the resonant frequency change of the biosensor was monitored in a wireless, magnetic manner. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 6.7 (a) Water contact angles for bare and surface-functionalized gold surfaces. (b) XPS S2p peaks at around 162 eV, indicating the formation of sulfur { gold bonded systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 6.8 AFM images (2 m 2 m) of the SAE10 phage on bare and surface- functionalized ME resonators: (a) reference, (b) bare gold (physical phage adsorption), (c) AC- (covalent phage attachment), (d) ALD-, and (e) MT- functionalized ME resonators. The white lines on the photographs were the paths from which the height pro les were measured. (f) The surface phage coverage values were 46.8%, 49.4%, 4.2%, and 5.2% for bare, AC-, ALD-, and MT-functionalized ME resonators, respectively. . . . . . . . . 132 6.9 Water contact angles for bare and AC-functionalized gold surfaces loaded with/without phages. The surfaces were washed with a 100-mM HEPES bu er containing various concentrations of Tween 20 (0 to 5 % v/v) and nally with DI water. The e ect of washing was found to be large on the immobilization stability of the physically adsorbed phages. . . . . . . . . 134 xvii 6.10 SEM images showing SA beads captured by the SAE10 phage on bare and surface-functionalized ME biosensors (all BSA blocked): (a) bare, (b) AC-, (c) ALD-, and (d) MT-functionalized ME biosensors. Non-speci c adsorption of SA beads was greatly reduced by BSA blocking: (e) bare and (f) AC-functionalized ME biosensors without phage. . . . . . . . . . 136 6.11 Dose-response plots showing the comparable performance of (a) bare and (b) AC-functionalized ME biosensors (1 mm 0.2 mm 15 m). . . . . 137 xviii List of Tables 1.1 Major performance criteria for biosensors. . . . . . . . . . . . . . . . . . 4 1.2 Infectious doses and incubation periods for the target pathogenic bacteria. 6 1.3 Recent Salmonella outbreaks in various food products in the United States. 7 1.4 Types of anthrax infection and associated fatality rates. . . . . . . . . . 9 2.1 Examples of culture-based bacterial detection. . . . . . . . . . . . . . . . 15 2.2 Examples of ELISA-based bacterial detection. . . . . . . . . . . . . . . . 17 2.3 Examples of PCR-based bacterial detection. . . . . . . . . . . . . . . . . 19 2.4 Comparison of the performance of major bacterial detection methods. . . 24 2.5 Variables and their values used for the calculations of the probability of detection. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.6 Lethal doses of B. anthracis spores in humans. . . . . . . . . . . . . . . 28 2.7 Minimum detectable number of pathogens for PCR- and biosensor-based methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3.1 Longevity of a landscape phage and monoclonal antibody at various tem- peratures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 3.2 Comparison between landscape phages and antibodies in terms of selec- tivity and production cost. . . . . . . . . . . . . . . . . . . . . . . . . . . 38 3.3 Phage clones used in this research. . . . . . . . . . . . . . . . . . . . . . 42 3.4 Materials properties for Metglas 2826MB and Fe80B20. . . . . . . . . . . 48 3.5 Sputtering conditions used for the fabrication of micron-scale sensor plat- forms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 3.6 Material costs for the sputtering targets. . . . . . . . . . . . . . . . . . . 54 xix 3.7 Di erently sized ME biosensors and their theoretical detection limits. . . 58 4.1 Mean resonant frequencies for the di erently sized biosensors. . . . . . . 74 4.2 LODs of the di erently sized biosensors for the adaxial and abaxial sur- faces of spinach leaves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 4.3 Surface geometric parameters for the adaxial and abaxial surfaces of spinach leaves. The values are the averages of 15 samples. . . . . . . . . . . . . . 81 4.4 Required number of biosensors, m, to obtain desired LODs for various values of P(D) (0.1, 0.5, and 0.8). . . . . . . . . . . . . . . . . . . . . . . 89 5.1 Number of bound spores and corresponding resonant frequency changes for measurement and control sensors. . . . . . . . . . . . . . . . . . . . . 111 6.1 Resonant frequency changes in Hz. . . . . . . . . . . . . . . . . . . . . . 118 6.2 Materials constants used in FE simulations. . . . . . . . . . . . . . . . . 120 xx List of Abbreviations AFM Atomic force microscopy AUDFS the Auburn University Detection and Food Safety Center BSA Bovine serum albumin CDC the Centers for Disease Control and Prevention DI Deionized DMF dimethylformamide DNA Deoxyribonucleic acid dNTP Deoxynucleotide triphosphate ELISA Enzyme-linked immunosorbent assay FE Finite element LOD Limit of detection ME Magnetoelastic PCR Polymerase chain reaction PDMS Poly(dimethylsiloxane) RNA Ribonucleic acid SAM Self-assembled monolayer SEM Scanning electron microscopy xxi TBS Tris-bu ered saline VBNC Viable but non-culturable XPS X-ray photoelectron spectroscopy xxii Chapter 1 Introduction 1.1 Background and need Food is essential for each individual to grow and stay healthy in daily life. How- ever, the past decades have been marked by a global increase in the outbreaks of food poisoning and associated illnesses. These public health problems are caused by the accidental supply and consumption of contaminated food, largely due to im- proper safety knowledge, perspectives, and practices of food producers [1] as well as insu cient consumer awareness [2]. Although substantial progress on food safety regulations has been made worldwide [3], up to 30% of the population even in indus- trialized countries su er from foodborne illnesses each year [4]. In the United States, for example, approximately 48 million cases of foodborne illnesses are estimated to occur annually, resulting in 128,000 hospitalizations, 3,000 deaths, and a3651.0 to a3677.7 billion economic losses [5{7]. At present, 31 foodborne pathogens, including bacteria, parasites, and viruses, are identi ed in the United States [7]. According to the Centers for Disease Control and Prevention (CDC) [5], most foodborne illnesses in the United States are caused by norovirus (58%), followed by nontyphoidal Salmonella spp. (11%), Clostridium per- fringens (10%), and Campylobacter spp (9%). In addition, the leading cause of both hospitalization and death is nontyphoidal Salmonella spp. (35% and 28%, respec- tively), which cause Salmonellosis, a major foodborne disease in most countries [4]. Food can be contaminated by these identi ed as well as unidenti ed pathogens at any stages of the supply chain (e.g., production, packaging, transportation, and retail). As a result, the food industry, the main party that is concerned with the presence 1 Figure 1.1: Economic cost - bene t assessment. of foodborn pathogens, is responsible for controlling the quality of food products. However, zero risk for all food products is unlikely to be achievable. Hence, one of the biggest challenges is to put e ective controls in place without unnecessarily in- creasing costs. In other words, the optimal level of food safety must be determined through economic cost - bene t assessments [8]. Figure 1.1 shows a typical represen- tation of the relationship between implicit price per unit of safety and level of safety. As can be seen from the upward sloping line of marginal social cost, it is inexpen- sive to improve safety at low levels, but further improvements are more costly. By contrast, marginal social bene t (i.e., society?s additional willingness to pay to avoid ill-health and the costs of treating ill-health) decreases as the level of safety increases, represented by the downward sloping line. At point A, the amount the society is willing to pay exceeds the amount it would cost to improve safety, indicating that it is worth allocating resources to produce more safety. By contrast, costs exceed ben- e ts at point B, meaning that too many resources are being devoted to safety. Only at point Qm are costs equal to bene ts per unit of safety, representing that e cient resource allocation occurs [8]. Hence, this point Qm is the sought safety level (Pm is 2 the corresponding price per unit of safety). From the above example of assessment, it is understandable that cost is an important factor that cannot be disregarded for the management of food safety risks. Foodborne illness is not the only problem that poses a severe risk to public safety. Since the 2001 anthrax mail attacks in the United States, bioterrorism, which makes use of bacteria, viruses, fungi, and/or toxins as a bioweapon, has been publicly recognized as an emerging danger. The CDC has, thus far, identi ed 35 potential bioterrorism agents and classi ed them into three categories [9]. For example, Bacillus anthracis, the etiologic agent of anthrax, is listed among the high-priority Category A agents, which have the potential for major public health impact. Comprehensive attempts to control these deadly biological agents have been made internationally by prohibiting their use and proliferation since the war era in the last century [10]. However, as is evident from the recent anthrax attacks, the attempts have not been Figure 1.2: Civilian biodefense funding by scal year, FY2001 - FY2013 (in a36millions). 3 entirely successful. As a result, the government of the United States has been en- hancing national biosecurity. In fact, the funding for civilian biodefense dramatically increased after the 2001 anthrax attacks, and over a364 billion of funding has been maintained since the scal year of 2002 as shown in Fig. 1.2 [11]. The budget transi- tion clearly indicates that there is a need for comprehensive biodefense systems that enable the nationwide surveillance and prevention of bioterrorism. In addition, a por- tion of the funding is dedicated to food defense, including the prevention of deliberate food contamination with pathogenic agents. As part of the ongoing e orts to secure food safety as well as to guard against possible bioterrorism, the role of pathogen detection technologies has become vi- tal. However, conventional and standard detection methods, including culture-, immunology-, and polymerase chain reaction-based methods, are generally expen- sive, time-consuming, and labor-intensive [3]. Hence, much research has been recently focused on developing label-free biosensors, which are meant to be low-cost, rapid, Table 1.1: Major performance criteria for biosensors. Criterion Description Sensitivity Ability to detect a small amount of pathogens in a reasonably small sample volume Selectivity Ability to distinguish among pathogens Assay time Short for a single test Thermal stability Ability to function at a wide range of temperatures Longevity Ability to retain detection capabilities for a fair period of time Assay protocol No reagent addition needed Measurement Direct and without pre-enrichment Format Highly automated format Operator No expertise needed Cost Inexpensive Size Compact and portable for on-site detection 4 and user-friendly, adequate for on-site pathogen detection for both food safety and biosecurity. Table 1.1 lists the major performance criteria for biosensors (partially adapted from [12]). Although some existing biosensors possess excellent performance, meeting various performance criteria simultaneously still remains a challenge. Hence, further research and development are essential before biosensors become a reliable, alternative solution. 1.2 Research objectives Magnetoelastic (ME) biosensors, a novel class of wireless, mass-sensitive biosen- sors, are among potential candidates that could overcome the above-mentioned perfor- mance challenge. In recent years, the Auburn University Detection and Food Safety Center (AUDFS) has begun research into the detection of pathogenic bacteria us- ing freestanding, strip-shaped ME biosensors combined with a landscape phage (i.e., genetically engineered phage) [13] as the biomolecular-recognition element [14{17]. These phage-based ME biosensors are not only rapid and thermally robust [16], but their wireless nature of detection o ers great exibility in design and use, which fa- cilitates on-site bacterial detection. In addition, the sensitivity of ME biosensors can be improved by reducing their dimensions [18], and the fabrication cost per sensor can be reduced via batch fabrication. Hence, the primary objectives of this research are (1) to further improve the cost-e ectiveness, rapidness, and sensitivity of the phage-based ME biosensors and (2) to demonstrate enhanced detection of pathogenic bacteria (i.e., Salmonella Typhimurium and Bacillus anthracis spores). In order to improve both cost-e ectiveness and sensitivity, micron- to millimeter-scale ME biosen- sors were batch-fabricated and used. In addition, the following two methodologies were employed to shorten assay time: 5 1. Direct detection of S. Typhimurium on fresh spinach leaves without any pre-test sample preparation (i.e., collection and puri cation of Salmonella-containing samples, followed by enrichment) 2. Detection of B. anthracis spores with the aid of a designed micro uidic ow cell, which ensures e cient physical contact between a biosensor and owing spores. Additionally, as potential ways to further enhance the detection capabilities of phage- based ME biosensors, the following e ects were studied: 1. E ects of mass position on the sensitivity of ME biosensors 2. E ects of surface functionalization of ME biosensors on surface phage coverage. 1.3 Target pathogenic bacteria to be detected S. Typhimurium and B. anthracis spores are target pathogenic bacteria to be detected in this research. Their median infectious doses and incubation periods are summarized in Table 1.2. Table 1.2: Infectious doses and incubation periods for the target pathogenic bacteria. Target pathogen Infectious dose Incubation period Ref. S. Typhimurium 100 to 1,000 cells (ingestion) 6 to 72 hr [19] B. anthracis 8,000 to 50,000 spores (inhalation) < 7 days [20] 1.3.1 Salmonella Typhimurium Nontyphoidal Salmonella spp. are important foodborne pathogens that cause gastroenteritis, bacteremia, and subsequent focal infection [21]. They are responsible for 11% of all foodborn illnesses in the United States, resulting in roughly 20,000 hospitalizations and 400 deaths each year [5]. Salmonellosis, caused by the ingestion 6 Table 1.3: Recent Salmonella outbreaks in various food products in the United States. Source Cause(s) Cases Hospitalizations Year Ref. Tomatoes S. Typhimurium 183 22 2006 [24] Peanut butter S. Tennessee 425 71 2007 [25] Cantaloupes S. Litch eld 51 > 16 2008 [26] Jalape~no peppers S. Saintpaul 1,442 > 286 2008 [27] Peanut butter S. Typhimurium 714 170 2009 [28] Alfalfa sprouts S. Saintpaul 235 7 2009 [29] Shell eggs S. Enteritidis 1,939 N/A 2010 [30] Cantaloupes S. Panama 20 3 2011 [31] Sprouts S. Enteritidis 25 3 2011 [32] Ground turkey S. Heidelberg 136 37 2011 [33] Chicken livers S. Heidelberg 190 30 2011 [34] Ground beef S. Typhimurium 20 8 2011 [35] Ground tuna S. Bareilly & S. Nchanga 425 55 2012 [36] Ground beef S. Enteritidis 46 12 2012 [37] Cantaloupes S. Typhimurium & S. Newport 270 101 2012 [38] Live poultry S. Montevideo 76 17 2012 [39] Mangoes S. Braenderup 121 25 2012 [40] Peanut butter S. Bredeney 30 4 2012 [41] of nontyphoidal Salmonella spp., is a major foodborne disease in most countries today [4] and usually contracted from various sources [22], including eggs, meat, poultry, and fresh produce as shown in Table 1.3. Among over 2,500 serovars capable of infecting humans and animals, S. Typhimurium is becoming one of the most prevalent serovars [23]. Hence, this pathogenic bacterium has been selected as one of the target pathogens to be detected in this research. Figure 1.3 shows a scanning electron micrograph of S. Typhimurium cells on a spinach leaf surface. 7 Figure 1.3: S. Typhimurium cells on a spinach leaf surface. 1.3.2 Bacillus anthracis spores B. anthracis, the etiologic agent of anthrax, is a rod-shaped, spore-forming bac- terium [42, 43]. Due to it?s ability to form a resistant spore (i.e., dehydrated, thick- walled cell), this bacterium can populate a wide range of environments, including soil, bodies of water and animal hosts [42]. Although B. anthracis spores are metabolically dormant, they can germinate and grow to a vast number of vegetative cells once en- tering a nutrient-rich host, which in turn causes the disease anthrax with high fatality rates (Table 1.4). Hence, the potential use of B. anthracis spores as a bioweapon is a signi cant public safety concern. When used in an aerosolized form, they could enter the bodies of individuals through inhalation. For example, the recent anthrax attacks that occurred in the United States in 2001 resulted in 11 cases of inhalational anthrax, 8 Table 1.4: Types of anthrax infection and associated fatality rates. Type Fatality rate Ref. Inhalational As high as 90% (< 50% with appropriate treatment) [44{46] Cutaneous 20% (< 1% with appropriate treatment) [45,46] Gastrointestinal 25 to 60% [45,46] 5 of whom died. Although early, proper antibiotic treatments have proven e ective in reducing the high fatality rates, such treatments are often di cult to provide due to initial, non-speci c symptoms in infected patients [44]. Hence, anthrax infection must be prevented through early detection of B. anthracis spores. Figure 1.4 shows a scanning electron micrograph of B. anthracis spores on a gold surface. Figure 1.4: B. anthracis spores on a gold surface. 9 1.4 Organization of this dissertation In this chapter, the need for high-performance biosensors for on-site pathogen detection was described, and the objectives of the present research were stated. The rest of this dissertation is organized as follows: Chapter 2 brie y reviews major bacterial detection methods and discusses rea- sons for the current shift towards the development of label-free biosensors. Chapter 3 describes the fundamentals, detection principle, and fabrication meth- ods of phage-based ME biosensors in depth. Chapter 4 presents an investigation into rapid, direct detection of S. Typhimurium on fresh spinach leaves. Various e ects, including the topography of spinach leaf sur- faces, the distribution of S. Typhimurium cells, and the size and number of ME biosensors, on the limit of detection will also be discussed. Chapter 5 presents an investigation into rapid, sensitive detection of B. anthracis spores using micron-scale ME biosensors in combination with a designed micro uidic ow cell. Chapter 6 investigates the e ects of mass position on the sensitivity of ME biosen- sors. Experimental and numerical results will be rst compared, and then, a formula predicting the sensor response for a single point-mass will be derived. In addition, the e ects of surface functionalization on surface phage coverage will be studied. Based on the results of these investigations, a concept of the patterning of the phage layer onto desired parts of the sensor surface will be introduced. Finally, Chapter 7 presents an overall summary and conclusions of this disserta- tion. 10 Bibliography [1] M. L. L. Ivey, J. T. LeJeune, and S. A. Miller, \Vegetable producers perceptions of food safety hazards in the midwestern USA," Food Control, vol. 26, no. 2, pp. 453 { 465, 2012. [2] C. Losasso, V. Cibin, V. Cappa, A. Roccato, A. Vanzo, I. Andrighetto, and A. Ricci, \Food safety and nutrition: Improving consumer behaviour," Food Control, vol. 26, no. 2, pp. 252 { 258, 2012. [3] V. Velusamy, K. Arshak, O. Korostynska, K. Oliwa, and C. Adley, \An overview of foodborne pathogen detection: In the perspective of biosensors," Biotechnology Advances, vol. 28, no. 2, pp. 232 { 254, 2010. [4] WHO, \Food safety and foodborne illness," 2007. fact sheet no. 237. [5] E. Scallan, R. M. Hoekstra, F. J. Angulo, R. V. Tauxe, M.-A. Widdowson, S. L. Roy, J. L. Jones, and P. M. Gri n, \Foodborne illness acquired in the United States - major pathogens," Emerging Infectious Diseases, vol. 17, no. 1, pp. 7 { 15, 2011. [6] E. Scallan, P. M. Gri n, F. J. Angulo, R. V. Tauxe, and R. M. Hoekstra, \Foodborne illness acquired in the United States - unspeci ed agents," Emerging Infectious Diseases, vol. 17, no. 1, pp. 16 { 22, 2011. [7] R. L. Schar , \Economic burden from health losses due to foodborne illness in the United States," Journal of food protection, vol. 75, no. 1, pp. 123 { 131, 2012. [8] W. B. Traill and A. Koenig, \Economic assessment of food safety standards: Costs and bene ts of alternative approaches," Food Control, vol. 21, no. 12, pp. 1611 { 1619, 2010. [9] M. Shannon, \Management of infectious agents of bioterrorism," Clinical Pedi- atric Emergency Medicine, vol. 5, no. 1, pp. 63 { 71, 2004. [10] D. P. Fidler, \Facing the global challenges posed by biological weapons," Mi- crobes and Infection, vol. 1, no. 12, pp. 1059 { 1066, 1999. [11] C. Franco and T. K. Sell, \Federal agency biodefense funding, FY2012 - FY2013," Biosecurity and Bioterrorism: Biodefense Strategy, Practice, and Sci- ence, vol. 10, no. 2, pp. 162 { 181, 2012. 11 [12] D. Ivnitski, I. Abdel-Hamid, P. Atanasov, and E. Wilkins, \Biosensors for de- tection of pathogenic bacteria," Biosensors and Bioelectronics, vol. 14, no. 7, pp. 599 { 624, 1999. [13] V. A. Petrenko, \Landscape phage as a molecular recognition interface for de- tection devices," Microelectronics Journal, vol. 39, no. 2, pp. 202 { 207, 2008. [14] S. Huang, H. Yang, R. Lakshmanan, M. Johnson, J. Wan, I.-H. Chen, H. W. III, V. Petrenko, J. Barbaree, and B. Chin, \Sequential detection of Salmonella typhimurium and Bacillus anthracis spores using magnetoelastic biosensors," Biosensors and Bioelectronics, vol. 24, no. 6, pp. 1730 { 1736, 2009. [15] R. S. Lakshmanan, R. Guntupalli, J. Hu, D.-J. Kim, V. A. Petrenko, J. M. Barbaree, and B. A. Chin, \Phage immobilized magnetoelastic sensor for the de- tection of Salmonella typhimurium," Journal of Microbiological Methods, vol. 71, no. 1, pp. 55 { 60, 2007. [16] J. Wan, H. Shu, S. Huang, B. Fiebor, I.-H. Chen, V. A. Petrenko, and B. A. Chin, \Phage-based magnetoelastic wireless biosensors for detecting Bacillus anthracis spores," IEEE Sensors Journal, vol. 7, pp. 470 { 477, 2007. [17] J. Wan, M. L. Johnson, R. Guntupalli, V. A. Petrenko, and B. A. Chin, \De- tection of Bacillus anthracis spores in liquid using phage-based magnetoelastic micro-resonators," Sensors and Actuators B: Chemical, vol. 127, no. 2, pp. 559 { 566, 2007. [18] M. L. Johnson, J. Wan, S. Huang, Z. Cheng, V. A. Petrenko, D.-J. Kim, I.- H. Chen, J. M. Barbaree, J. W. Hong, and B. A. Chin, \A wireless biosensor using microfabricated phage-interfaced magnetoelastic particles," Sensors and Actuators A: Physical, vol. 144, no. 1, pp. 38 { 47, 2008. [19] http://www.msdsonline.com/resources/msds-resources/ free-safety-data-sheet-index/salmonella-spp.aspx. [20] http://www.msdsonline.com/resources/msds-resources/ free-safety-data-sheet-index/bacillus-anthracis.aspx. [21] D. Acheson and E. L. Hohmann, \Nontyphoidal Salmonellosis," Clinical Infec- tious Diseases, vol. 32, no. 2, pp. 263 { 269, 2001. [22] http://www.harrisonspractice.com/practice/ub/view/Harrisons%20Practice/ 141052/all/Nontyphoidal Salmonellosis. [23] http://www.safe-poultry.com/ParatyphoidSalmonella.asp. [24] http://www.cdc.gov/ncidod/dbmd/diseaseinfo/salmonellosis 2006/ 110306 outbreak notice.htm. 12 [25] http://www.cdc.gov/ncidod/dbmd/diseaseinfo/salmonellosis 2007/ 030707 outbreak notice.htm. [26] http://www.cdc.gov/salmonella/litch eld/. [27] http://www.cdc.gov/salmonella/saintpaul/jalapeno/index.html. [28] http://www.cdc.gov/salmonella/typhimurium/update.html. [29] http://www.cdc.gov/salmonella/saintpaul/alfalfa/. [30] http://www.cdc.gov/salmonella/enteritidis/index.html. [31] http://www.cdc.gov/salmonella/panama0311/062311/index.html. [32] http://www.cdc.gov/salmonella/sprouts-enteritidis0611/070611/index.html. [33] http://www.cdc.gov/salmonella/heidelberg/111011/index.html. [34] http://www.cdc.gov/salmonella/heidelberg-chickenlivers/011112/index.html. [35] http://www.cdc.gov/salmonella/typhimurium-groundbeef/020112/index.html. [36] http://www.cdc.gov/salmonella/bareilly-04-12/index.html. [37] http://www.cdc.gov/salmonella/enteritidis-07-12/index.html. [38] http://www.cdc.gov/salmonella/typhimurium-cantaloupe-08-12/index.html. [39] http://www.cdc.gov/salmonella/montevideo-06-12/index.html. [40] http://www.cdc.gov/salmonella/braenderup-08-12/index.html. [41] http://www.cdc.gov/salmonella/bredeney-09-12/index.html. [42] A. Driks, \The Bacillus anthracis spore," Molecular Aspects of Medicine, vol. 30, no. 6, pp. 368 { 373, 2009. [43] T. M. Koehler, \Bacillus anthracis physiology and genetics," Molecular Aspects of Medicine, vol. 30, no. 6, pp. 386 { 396, 2009. [44] T. V. Inglesby, T. O?Toole, D. A. Henderson, J. G. Bartlett, M. S. Ascher, E. Eitzen, A. M. Friedlander, J. Gerberding, J. Hauer, J. Hughes, J. McDade, M. T. Osterholm, G. Parker, T. M. Perl, P. K. Russell, and K. Tonat, \Anthrax as a biological weapon, 2002: Updated recommendations for management," The Journal of the American Medical Association, vol. 287, no. 17, pp. 2236 { 2252, 2002. [45] http://www.bt.cdc.gov/agent/anthrax/faq/. [46] http://www.upmc-biosecurity.org/website/our work/ biological-threats-and-epidemics/fact sheets/anthrax.html. 13 Chapter 2 Review of the Literature on Bacterial Detection Methods This chapter reviews major bacterial detection methods and discusses reasons for the current shift towards the development of label-free biosensors. 2.1 Conventional detection methods Conventional methods for the detection of pathogenic bacteria rely on speci c microbiological or biochemical identi cation. Three major conventional methods are culture-, immunology-, and polymerase chain reaction-based methods as shown in Fig. 2.1. Although these methods can be highly sensitive, selective, and reliable, their application to on-site bacterial detection is greatly restricted by several drawbacks, including long assay times, high cost, and cumbersome procedures, requiring trained personnel. Figure 2.1: Conventional methods for bacterial detection. 2.1.1 Culture-based methods Culture-based methods remain the most reliable and commonly used techniques for bacterial detection. These methods are capable of identifying a small number of 14 pathogenic bacteria (down to single bacteria). However, cumbersome assay steps, including pre-enrichment, selective enrichment, colony counting, biochemical screen- ing, and serological con rmation, are generally required [1]. As a result, depending on bacterial species and/or strains, these culture-based methods may take days to weeks to yield results, which hinders their use in on-site bacterial detection. In addi- tion, some viable bacteria in the environment may enter a dormant state and become non-culturable (i.e., viable but non-culturable (VBNC) state), which leads to an un- derestimation of the quantity of the bacteria or a failure to identify the bacteria in a contaminated sample [2]. Table 2.1 shows examples of culture-based bacterial detection and their assay times. Table 2.1: Examples of culture-based bacterial detection. Detected pathogen Assay time Ref. Escherichia coli O157:H7 2 days [3] Salmonella Enteritidis 4 to 8 days [4] Listeria monocytogenes up to 7 days [5,6] Campylobacter fetus 14 to 16 days [7] 2.1.2 Immunology-based methods Immunology-based methods, the majority of which rely on an antibody - antigen binding, have been widely used for the detection of pathogenic bacteria, including Escherichia coli, Salmonella spp., Listeria monocytogenes, Campylobacter spp., and Staphylococcal enterotoxins [2]. Among existing methods, enzyme-linked immunosor- bent assay (ELISA), which is relatively rapid and versatile, is the most commonly used technique. Figure 2.2 illustrates a typical procedure for sandwich ELISA, a commonly used ELISA variant. The assay steps are as follows: 15 Figure 2.2: Typical procedure for sandwich ELISA. 1. Immobilize a capture antibody on the surface of each well of a microtiter plate (often called an ELISA plate). Then, wash the plate so that any unbound antibodies are removed. 2. Block any non-speci c adsorption sites on the well surface with a surface block- ing agent (usually, bovine serum albumin or casein). 3. Apply a sample that contains a target antigen to the plate and allow the capture antibody to bind with the antigen. Then, wash the plate to remove any unbound antigens. 4. Add a primary antibody and allow it to bind with the antigen. Then, wash the plate. 5. Add an enzyme-linked secondary antibody that binds with the primary anti- body. Then, wash the plate. 16 6. Finally, add a substrate that can be converted by the enzyme into a color or electrochemical signal for measurement. Although ELISA-based methods are much more rapid than culture-based methods, hours to days are still required to yield results [8{10]. In addition, their limits of detec- tion (LODs) are not competitive with those of culture-based methods. Furthermore, cumbersome assay procedures (i.e., a series of washing and addition of reagents) make ELISA-based methods unsuitable for on-site bacterial detection. Table 2.2 shows ex- amples of ELISA-based bacterial detection and their LODs. Table 2.2: Examples of ELISA-based bacterial detection. Detected pathogen LOD Ref. Escherichia coli O157 103 to 104 cfu/ml [11,12] Salmonella serovars 106 cells/ml [13] Listeria monocytogenes 103 cells/ml [14] Campylobacter fetus 105 cells/ml [7] 2.1.3 Polymerase chain reaction-based methods The polymerase chain reaction (PCR) is a biochemical technique to produce mil- lions of copies of a fragment of a nucleic acid (usually, deoxyribonucleic acid (DNA), which is more stable than ribonucleic acid (RNA)). PCR-based methods have been widely used to identify or detect pathogenic bacteria, including Salmonella aureus, Listeria monocytogenes, Bacillus cereus, Escherichia coli O157:H7, and Campylobac- ter jejuni [2]. As shown in Fig. 2.3, the PCR typically requires a series of 20 to 40 thermal cycles. In each cycle, there are three discrete temperature steps as described below: 1. Denaturation: separating double-stranded DNA into a pair of single-stranded DNA templates at a temperature of around 95 C. 17 Figure 2.3: Schematic illustration of the PCR cycle. 2. Annealing: allowing annealing of forward and reverse primers to the single- stranded DNA templates at a temperature of 50 to 65 C. 3. Elongation: extending the primers with the aid of DNA polymerase to synthe- size complementary strands at a temperature of around 70 C. In this way, the number of amplicons doubles after each cycle, resulting in exponential ampli cation in the amount of the target DNA sequence. Since each cycle requires only several minutes, millions of amplicons can be produced within a few hours. After thermal cycling, the nal PCR products are typically analyzed by gel electrophoresis. PCR-based methods possess the capability of detecting a small amount of target DNA (down to a few DNA molecules [15, 16]) as well as o er high speci city and 18 accuracy. In addition, they are relatively rapid (i.e., a few hours of assay time) when compared with culture- and immunology-based methods. However, their use in on-site bacterial detection is restricted by a number of shortcomings. They require pure DNA samples and speci c primers for avoiding false ampli cation, expensive reagents (e.g., DNA polymerase, deoxynucleotide triphosphate (dNTP), and other additives), and hours of thermal cycles, followed by a gel electrophoresis-based analysis, which usually takes one additional hour. In other words, these PCR-based methods are complex, expensive, and still time-consuming. Although newer PCR variants, including real- time PCR [17], digital PCR [18], and micro uidic PCR [19], can o er a much shorter assay time with less volumes of reagents, the use of a uorescent-labeled DNA probe as well as an optical detector for the acquisition of uorescence signals is additionally needed, leading to an increase in cost and assay complexity. Furthermore, PCR-based methods cannot generally discriminate between viable and non-viable cells [20], as well as the extraction of DNA from a resistant bacterial spore, such as a B. anthracis spore, remains a challenge [21]. Examples of PCR-based bacterial detection and their LODs are shown in Table 2.3. Table 2.3: Examples of PCR-based bacterial detection. Detected pathogen LOD Ref. Escherichia coli 102 cells/ml [22] Salmonella Enteritidis 1 cfu/25 g or ml of food samples [23] Listeria monocytogenes 3 cfu/g in ground beef [24] Campylobacter spp. 100 to 150 cfu/ml [25] Legionella pneumophila < 10 cfu/ml [26] 19 2.2 Biosensors as promising bacterial detection methods Figure 2.4 shows the number of research articles published between 1985 and 2005 on di erent bacterial detection methods [20]. As can be seen, the most popular detection methods were PCR-, culture-, and ELISA-based methods, which is due to their low LODs and high reliability as mentioned in the previous sections. However, in addition to these conventional methods, emerging biosensor technologies have been drawing much attention in recent years. In fact, the global market for biosensors in 2012 is estimated to be a368.5 billion and projected to reach a3616.8 billion by 2018 [27]. Biosensor technologies come with promises of equally reliable results in much shorter times [20]. Figure 2.4: Number of research articles published between 1985 and 2005 on di erent bacterial detection methods. 2.2.1 De nition of a biosensor A biosensor is an analytical device that converts a biological response into an electrical signal. Two principal components of a biosensor are: (1) a biomolecular- recognition element, which recognizes and speci cally binds with a target analyte, 20 and (2) a signal transducer, which converts the recognition event into a measurable electrical signal. Figure 2.5 shows a schematic diagram of a biosensor. When a biomolecular- recognition event occurs, a signal can be instantaneously generated by the transducer. This initial, small input signal from the transducer is, then, ampli ed, processed, and sent to an output system for display or further analyses. Biosensors can be rapid, sensitive, target-speci c, and portable, which makes them suitable for use in a variety of elds, including medical care, environmental monitoring, food safety, and biosecurity. Biosensors can be classi ed by their recognition elements and/or signal transduction methods as shown in Fig. 2.6. Figure 2.5: Schematic diagram of a biosensor. 21 Figure 2.6: Classi cation of biosensors. 2.2.2 Biomolecular-recognition elements Biomolecular-recognition elements are responsible for speci cally binding a tar- get analyte to a biosensor. They are generally immobilized on the surface of a sig- nal transducer. Antibodies, nucleic acids, and enzymes are three major types of biomolecular-recognition elements as can be seen in Fig. 2.6. In recent years, how- ever, a number of attempts have been made in employing landscape phages (i.e., genetically engineered phages) as the biomolecular-recognition element for the detec- tion of pathogenic bacteria [28{31]. Landscape phages are highly tailorable, speci c, relatively inexpensive, and thermally robust (much better than commonly used anti- bodies) [32]. Hence, in this research, landscape phages were a nity-selected for the target pathogenic bacteria (i.e., S. Typhimurium and B. anthracis spores) and used. 22 2.2.3 Signal transducers Signal transducers are responsible for converting a biomolecular-recognition event into a measurable electrical signal. In the past decades, a wide variety of transduction methods has been developed. Among them, optical, electrochemical, and mass-based methods are the most commonly used methods (Fig. 2.6). Each of these major types of transducers contains many di erent subtypes. In addition, they can be further classi ed into labeled and label-free methods. While the labeled methods depend on the detection of a speci c label (e.g., uorescent, chemiluminescent, and radioactive labels), the label-free methods are based on the direct measurement of a phenomenon occurring on a transducer surface [2]. In this research, freestanding, strip-shaped mag- netoelastic (ME) transducers, a novel class of mass-based transducers, were combined with the above-mentioned landscape phages and used. Compared with conventional mass-based transducers (e.g., piezoelectric transducers), ME transducers possess ad- vantageous features, such as low-cost production, wireless signal transduction, and thus, exibility in biosensor design. These unique characteristics of ME transducers facilitate on-site bacterial detection. 2.3 Conventional detection methods vs. biosensors Table 2.4 compares the LODs and assay times of major bacterial detection meth- ods. As mentioned earlier, the conventional detection methods possess low LODs, and even single pathogenic bacteria can be detected with culture-based methods. How- ever, the conventional methods are all time-consuming (i.e., up to weeks of assay time required), and thus, the nal results can not be obtained rapidly. By contrast, biosensors are generally much more rapid (i.e., on the order of minutes and up to hours of assay time), which is a distinct advantage for on-site bacterial detection for food safety and biosecurity. However, their LODs need to be improved so that they can be competitive with the conventional detection methods. In addition, other 23 Table 2.4: Comparison of the performance of major bacterial detection methods. Detection method LOD Assay time Ref. Conventional methods Culture Down to a single cell Days to weeks [3{7] ELISA 103 to 106 cfu/ml Hours to days [7,11{14] PCR Down to a single cell Hours [16,22{26] Optical biosensors Surface plasmon resonance 50 to 105 cfu/ml 15 min to hours [33{36] Resonant mirror 103 spores 10 min [37] Interferometer 5 106 cfu/ml < 40 min [38] Ring resonator 105 cfu/ml < 60 min [39] Bioluminescence 10 cfu/ml 20 min [40] Electrochemical biosensors Amperometric 101 to 103 cfu/ml Minutes to hours [41{45] Potentiometric 101 to 103 cfu/ml 30 min to 1.5 hr [46{48] Impedimetric 2 cfu/ml 45 min [49] Conductometric 61 cfu/ml 8 min [50] Mass-based biosensors Quartz-crystal microbalance 103 spores/ml < 30 min [51] Love-wave < 200 spores/ml 5 min [52] Silicon cantilevers 1 spore in air hours [53] 50 spores in water hours [53] Piezoelectric cantilevers 300 spores/ml < 20 min [54] Magnetoelastic cantilevers 105 cfu/ml < 120 min [55] Magnetoelastic strips 103 cfu/ml < 30 min [56] performance criteria previously shown in Table 1.1 need to be met before biosensors become reliable alternatives. 24 2.3.1 Probability of detection: PCR vs. biosensors Although there is still room for improvement, biosensors are promising tools for pathogen detection, which may be explained with a statistical model reported by Sabelnikov et al [15]. The model can be used to estimate the probability of detection for an aerosolized pathogen (e.g., aerosolized B. anthracis spores) using a model detector or its network. A model detector is de ned as a single device that consists of an aerosol sampler and a detection device based on any of known bacterial detection methods, such as PCR-, immunology-, and biosensor-based methods. In addition, a network of model detectors consists of m single model detectors, each of which operates in the same way and deals with the same amount of a sample that contains a target pathogen. The assumptions for this statistical model are as follows [15]: 1. There is a space that contains aerosolized particles of a pathogen. These pathogen particles are distributed in the space according to a Poisson distri- bution with a parameter, , which is equal to the mean concentration of the pathogen. 2. The aerosol sampler intakes the air with a ow rate, Ws, and concentrates the aerosolized pathogen with an e ciency, Ke, into a liquid collective sample with a volume, Vc. 3. The time of sampling is set equal to the time of inhalation of the pathogen by an individual. Since the total number of the pathogens inhaled by the exposed individual, Di, is equal to the concentration of the pathogen in the air multiplied by the time of exposure and the inhalation rate, Wh, this assumption allows to exclude time and concentration factors from all calculations. 4. n individual samples of identical volume, Vs, are simultaneously tested for de- tection. 25 5. The pathogen can be detected in a single sample (i.e, n = 1) with a probability of 100% only if its amount in the sample is greater than or equal to a certain threshold value, I. In other words, the value I represents the LOD of the detection device (i.e., the minimum detectable number of pathogens per sample volume, Vs). For simpli cation, neither false positives nor false negatives are allowed. Based on the above assumptions, the probability of detection of the pathogen in a single sample (i.e., n = 1), Pds, may be expressed as Pds = 1 I 1X k=0 F(k); (2.1) F(k) = ( Vs) ke Vs k! ; (2.2) = KeDiWsW hVc ; (2.3) where F(k) is the probability of nding exactly k pathogens in the sample volume, Vs. In addition, the probability of detection with one model detector (i.e., m = 1 with n samples), Pdn, is equal to the probability that the pathogen can be detected in at least one of n individual samples, which can be calculated by Pdn = 1 (1 Pds)n: (2.4) By analogy, the probability of detection with a network of m model detectors, Pdm, can be computed by Pdm = 1 (1 Pdn)m: (2.5) The use of the above equations allows one to compare di erent bacterial detection methods in terms of their probability of detection with respect to the inhalation dose of a target pathogen. Here, as an example, comparisons will be made for PCR- 26 and biosensor-based methods. In all calculations, Ke and Vc were kept constant and equal to 0.8 and 10 ml, respectively, which are currently used in the most advanced commercial samplers [58, 59]. In addition, an inhalation rate (Wh) of 11 l/min for adult humans [60] was used. For convenience, all the variables and their values used are summarized in Table 2.5. Table 2.5: Variables and their values used for the calculations of the probability of detection. Variable Values Ref. Di: total number of inhaled pathogens 1 to 100 Ws: intake ow rate of the sampler 1,000 l/min [57,58] Ke: e cacy coe cient of the sampler (ratio of the number of concentrated pathogens to the number of sampled pathogens) 0.8 [58] Vc: volume of a liquid collective sample 10 ml [58,59] Vs: volume of an individual sample for detec- tion (1) 50 to 500 l for PCR (2) 1 ml for biosensors [15] n: number of individual samples (1) 96 or 384 for PCR (2) 1 for biosensors [15] I: detectable number of pathogens per sample volume, Vs (1) 15 for PCR (2) 50 for biosensors [15] m: number of model detectors (1) 1 for PCR (2) 1 to 4 for biosensors Wh: inhalation rate for adult humans 11 l/min [60] Figure 2.7 shows the results for PCR-based detectors. These detectors can si- multaneously test 96 or 384 samples (i.e., n = 96 or 384, based on the standard 96- or 384-well format of PCR technologies) with I = 15, which is of the best commercial, eld-operated PCR device [15]. In addition, various values of Vs (50, 100, and 500 l) were used in the calculations. For a standard PCR-based detector (n = 96 and Vs = 50 l), it was found that 36 and more pathogens can be detected with a probability of 100% (solid curve). In addition, a slightly better result was obtained with n = 384, which reduces the minimum detectable number of pathogens with a probability 27 Figure 2.7: Probability of detection with respect to the inhalation dose of the target pathogen for PCR-based detectors. The following values were used in the calculations: Ws = 1,000 l/min, Ke = 0.8, I = 15, and m = 1. of 100% to be 28 (dashed curve). However, for both detectors, detection of less than 10 pathogens was found to be hardly possible, which may lead to certain risks of lethality in humans for some existing pathogens. For instance, Table 2.6 shows lethal doses of B. anthracis spores in humans for di erent levels of lethality. The LD10, the lethal dose at which 10% of the population is expected to die, may be as low as 50 to 98 spores, and even lower lethal doses for the LD5 and LD1 have been reported. Comparing the above calculated and these reported lethal doses clearly indicates that both of the standard PCR-based detectors would fail to detect B. anthracis spores in doses that could still cause a lethality of 5%. Although an increase in the volume of Table 2.6: Lethal doses of B. anthracis spores in humans. Level of lethality Lethal dose [spores] Ref. LD10 50 to 98 [61] LD5 14 to 28 [61] LD1 1 to 3 [61] 28 individual samples, Vs, further reduces the minimum detectable number of pathogens (solid curves with squares and circles in Fig. 2.7), the use of such larger volumes per sample (i.e., Vs = 100 or 500 l) is not practical because it obviously increases costs due to a corresponding increase in the volumes of reagents required for the PCR (i.e., DNA polymerase, dNTPs, and other additives). Figure 2.8: Probability of detection with respect to the inhalation dose of the tar- get pathogen for biosensor-based detectors. The following values were used in the calculations: Ws = 1,000 l/min, Ke = 0.8, Vs = 1 ml, and n = 1. Figure 2.8 shows the results for biosensor-based detectors with various values of I (10, 50, and 100 pathogens per Vs). Biosensors generally deal with a much larger individual sample volume (1 ml) than standard PCR-based methods (50 l). Hence, if the value of I is su ciently small, they could outperform standard PCR-based detectors as shown in Table 2.7. It can be seen that the minimum detectable number of pathogens for the biosensor-based detectors dramatically decreases as the value of I is decreased. In addition, when a network of four detectors (i.e., m = 4) with I = 10 is used, down to four pathogens can be detected. Hence, these calculation results indicate that the use of biosensors can be a better choice than PCR-based detection 29 methods. Particularly, the use of a network of biosensors o ers a promise in rapid, sensitive detection of target pathogens. Table 2.7: Minimum detectable number of pathogens for PCR- and biosensor-based methods. Detection method I Vs n m Minimum detectable number of pathogens Standard PCR 15 50 l 96 1 36 Biosensor 100 1 ml 1 1 29 Biosensor 50 1 ml 1 1 19 Biosensor 10 1 ml 1 1 9 Biosensor 10 1 ml 1 4 4 30 Bibliography [1] J. J. Gooding, \Biosensor technology for detecting biological warfare agents: Recent progress and future trends," Analytica Chimica Acta, vol. 559, no. 2, pp. 137 { 151, 2006. [2] V. Velusamy, K. Arshak, O. Korostynska, K. Oliwa, and C. Adley, \An overview of foodborne pathogen detection: In the perspective of biosensors," Biotechnology Advances, vol. 28, no. 2, pp. 232 { 254, 2010. [3] G. Ngwa, R. Schop, S. Weir, C. Le on-Velarde, and J. Odumeru, \Detection and enumeration of E. coli O157:H7 in water samples by culture and molecular methods," Journal of Microbiological Methods, vol. 92, no. 2, pp. 164 { 172, 2013. [4] B. Lungu, W. D. Waltman, R. D. Berghaus, and C. L. Hofacre, \Comparison of a real-time PCR method with a culture method for the detection of Salmonella enterica serotype Enteritidis in naturally contaminated environmental samples from integrated poultry houses," Journal of food protection, vol. 75, no. 4, pp. 743 { 747, 2012. [5] E. de Boer and R. R. Beumer, \Methodology for detection and typing of food- borne microorganisms," International Journal of Food Microbiology, vol. 50, no. 12, pp. 119 { 130, 1999. [6] S. Artault, J. L. Blind, J. Delaval, Y. Dureuil, and N. Gaillard, \Detecting Listeria monocytogenes in food," International Food Hygiene, vol. 12, p. 23, 2001. [7] B. Brooks, J. Devenish, C. Lutze-Wallace, D. Milnes, R. Robertson, and G. Berlie-Surujballi, \Evaluation of a monoclonal antibody-based enzyme-linked immunosorbent assay for detection of Campylobacter fetus in bovine preputial washing and vaginal mucus samples," Veterinary Microbiology, vol. 103, no. 12, pp. 77 { 84, 2004. [8] K. D. King, J. M. Vanniere, J. L. LeBlanc, K. E. Bullock, and G. P. Anderson, \Automated ber optic biosensor for multiplexed immunoassays," Environmental Science & Technology, vol. 34, no. 13, pp. 2845 { 2850, 2000. [9] F. S. Ligler, G. P. Anderson, P. T. Davidson, R. J. Foch, J. T. Ives, K. D. King, G. Page, D. A. Stenger, and J. P. Whelan, \Remote sensing using an airborne biosensor," Environmental Science & Technology, vol. 32, no. 16, pp. 2461 { 2466, 1998. 31 [10] K. Tsuchiya and C. C. dUrsel, \Development of a sensitive ELISA using three monoclonal antibodies against lipopolysaccharides to detect Xanthomonas campestris pv. vesicatoria, the causal agent of bacterial spot of tomato and pep- per," Journal of General Plant Pathology, vol. 70, pp. 21 { 26, 2004. [11] B. W. Blais, J. Leggate, J. Bosley, and A. Martinez-Perez, \Comparison of uo- rogenic and chromogenic assay systems in the detection of Escherichia coli O157 by a novel polymyxin-based ELISA," Letters in Applied Microbiology, vol. 39, no. 6, pp. 516 { 522, 2004. [12] M. Feng, Q. Yong, W. Wang, H. Kuang, L. Wang, and C. Xu, \Development of a monoclonal antibody-based ELISA to detect Escherichia coli O157:H7," Food and Agricultural Immunology, pp. 1 { 7, 2012. [13] L. P. Mans eld and S. Forsythe, \The detection of Salmonella serovars from animal feed and raw chicken using a combined immunomagnetic separation and ELISA method," Food Microbiology, vol. 18, no. 4, pp. 361 { 366, 2001. [14] E. L. Crowley, C. K. O?Sullivan, and G. G. Guilbault, \Increasing the sensitivity of Listeria monocytogenes assays: Evaluation using ELISA and amperometric detection," Analyst, vol. 124, pp. 295 { 299, 1999. [15] A. Sabelnikov, V. Zhukov, and R. Kempf, \Probability of real-time detection versus probability of infection for aerosolized biowarfare agents: A model study," Biosensors and Bioelectronics, vol. 21, no. 11, pp. 2070 { 2077, 2006. [16] K. M. Hong, H. Najjar, M. Hawley, and R. D. Press, \Quantitative real-time PCR with automated sample preparation for diagnosis and monitoring of cy- tomegalovirus infection in bone marrow transplant patients," Clinical Chemistry, vol. 50, no. 5, pp. 846 { 856, 2004. [17] M. Kubista, J. M. Andrade, M. Bengtsson, A. Forootan, J. Jon ak, K. Lind, R. Sindelka, R. Sj oback, B. Sj ogreen, L. Str ombom, A. St ahlberg, and N. Zoric, \The real-time polymerase chain reaction," Molecular Aspects of Medicine, vol. 27, no. 23, pp. 95 { 125, 2006. [18] B. Vogelstein and K. W. Kinzler, \Digital PCR," Proceedings of the National Academy of Sciences, vol. 96, no. 16, pp. 9236 { 9241, 1999. [19] J. Liu, M. Enzelberger, and S. Quake, \A nanoliter rotary device for polymerase chain reaction," Electrophoresis, vol. 23, no. 10, pp. 1531 { 1536, 2002. [20] O. Lazcka, F. J. D. Campo, and F. X. M. noz, \Pathogen detection: A per- spective of traditional methods and biosensors," Biosensors and Bioelectronics, vol. 22, no. 7, pp. 1205 { 1217, 2007. [21] P. Belgrader, M. Okuzumi, F. Pourahmadi, D. A. Borkholder, and M. Northrup, \A micro uidic cartridge to prepare spores for PCR analysis," Biosensors and Bioelectronics, vol. 14, no. 1011, pp. 849 { 852, 2000. 32 [22] J. A. Higgins, S. Nasarabadi, J. S. Karns, D. R. Shelton, M. Cooper, A. Gbakima, and R. P. Koopman, \A handheld real time thermal cycler for bacterial pathogen detection," Biosensors and Bioelectronics, vol. 18, no. 9, pp. 1115 { 1123, 2003. [23] C. Almeida, L. Cerqueira, N. Azevedo, and M. Vieira, \Detection of Salmonella enterica serovar Enteritidis using real time PCR, immunocapture assay, PNA FISH and standard culture methods in di erent types of food samples," Inter- national Journal of Food Microbiology, vol. 161, no. 1, pp. 16 { 22, 2013. [24] P. G. Klein and V. K. Juneja, \Sensitive detection of viable Listeria monocyto- genes by reverse transcription-PCR.," Applied and Environmental Microbiology, vol. 63, no. 11, pp. 4441 { 4448, 1997. [25] M. Lund, S. Nordentoft, K. Pedersen, and M. Madsen, \Detection of Campy- lobacter spp. in chicken fecal samples by real-time PCR," Journal of Clinical Microbiology, vol. 42, no. 11, pp. 5125 { 5132, 2004. [26] P. Villari, E. Motti, C. Farullo, and I. Torre, \Comparison of conventional culture and PCR methods for the detection of Legionella pneumophila in water," Letters in Applied Microbiology, vol. 27, no. 2, pp. 106 { 110, 1998. [27] http://www.pharmaceutical-market-research.com/publications/ nance investment/biosensors global market overview.html. [28] S. Huang, H. Yang, R. Lakshmanan, M. Johnson, J. Wan, I.-H. Chen, H. W. III, V. Petrenko, J. Barbaree, and B. Chin, \Sequential detection of Salmonella typhimurium and Bacillus anthracis spores using magnetoelastic biosensors," Biosensors and Bioelectronics, vol. 24, no. 6, pp. 1730 { 1736, 2009. [29] M. L. Johnson, J. Wan, S. Huang, Z. Cheng, V. A. Petrenko, D.-J. Kim, I.- H. Chen, J. M. Barbaree, J. W. Hong, and B. A. Chin, \A wireless biosensor using microfabricated phage-interfaced magnetoelastic particles," Sensors and Actuators A: Physical, vol. 144, no. 1, pp. 38 { 47, 2008. [30] R. S. Lakshmanan, R. Guntupalli, J. Hu, D.-J. Kim, V. A. Petrenko, J. M. Barbaree, and B. A. Chin, \Phage immobilized magnetoelastic sensor for the de- tection of Salmonella typhimurium," Journal of Microbiological Methods, vol. 71, no. 1, pp. 55 { 60, 2007. [31] J. Wan, H. Shu, S. Huang, B. Fiebor, I.-H. Chen, V. A. Petrenko, and B. A. Chin, \Phage-based magnetoelastic wireless biosensors for detecting Bacillus anthracis spores," IEEE Sensors Journal, vol. 7, pp. 470 { 477, 2007. [32] V. A. Petrenko, \Landscape phage as a molecular recognition interface for de- tection devices," Microelectronics Journal, vol. 39, no. 2, pp. 202 { 207, 2008. [33] G. C. Bokken, R. J. Corbee, F. van Knapen, and A. A. Bergwer , \Immunochem- ical detection of Salmonella group B, D and E using an optical surface plasmon 33 resonance biosensor," FEMS Microbiology Letters, vol. 222, no. 1, pp. 75 { 82, 2003. [34] B.-K. Oh, Y.-K. Kim, K. W. Park, W. H. Lee, and J.-W. Choi, \Surface plasmon resonance immunosensor for the detection of Salmonella typhimurium," Biosen- sors and Bioelectronics, vol. 19, no. 11, pp. 1497 { 1504, 2004. [35] D. Zhang, Y. Yan, Q. Li, T. Yu, W. Cheng, L. Wang, H. Ju, and S. Ding, \Label- free and high-sensitive detection of Salmonella using a surface plasmon resonance DNA-based biosensor," Journal of Biotechnology, vol. 160, no. 34, pp. 123 { 128, 2012. [36] Y. Wang, W. Knoll, and J. Dostalek, \Bacterial pathogen surface plasmon reso- nance biosensor advanced by long range surface plasmons and magnetic nanopar- ticle assays," Analytical Chemistry, vol. 84, no. 19, pp. 8345 { 8350, 2012. [37] T. E. Love, C. Redmond, and C. N. Mayers, \Real time detection of anthrax spores using highly speci c anti-EA1 recombinant antibodies produced by com- petitive panning," Journal of Immunological Methods, vol. 334, no. 12, pp. 1 { 10, 2008. [38] B. H. Schneider, J. G. Edwards, and N. F. Hartman, \Hartman interferometer: versatile integrated optic sensor for label-free, real-time quanti cation of nucleic acids, proteins, and pathogens," Clinical Chemistry, vol. 43, no. 9, pp. 1757 { 1763, 1997. [39] A. Ramachandran, S. Wang, J. Clarke, S. Ja, D. Goad, L. Wald, E. Flood, E. Knobbe, J. Hryniewicz, S. Chu, D. Gill, W. Chen, O. King, and B. Little, \A universal biosensing platform based on optical micro-ring resonators," Biosensors and Bioelectronics, vol. 23, no. 7, pp. 939 { 944, 2008. [40] J. Qiu, Y. Zhou, H. Chen, and J.-M. Lin, \Immunomagnetic separation and rapid detection of bacteria using bioluminescence and micro uidics," Talanta, vol. 79, no. 3, pp. 787 { 795, 2009. [41] Z. Muhammad-Tahir and E. Alocilja, \Fabrication of a disposable biosensor for Escherichia coli O157:H7 detection," Sensors Journal, IEEE, vol. 3, pp. 345 { 351, Aug. 2003. [42] Z. Muhammad-Tahir and E. Alocilja, \A disposable biosensor for pathogen de- tection in fresh produce samples," Biosystems Engineering, vol. 88, no. 2, pp. 145 { 151, 2004. [43] C. Ruan, H. Wang, and Y. Li, \A bienzyme electrochemical biosensor coupled with immunomagnetic separation for rapid detection of Escherichia coli O157:H7 in food samples," Transactions of the ASABE, vol. 45, no. 1, pp. 249 { 255, 2002. 34 [44] L. Yang, C. Ruan, and Y. LI, \Rapid detection of Salmonella typhimurium in food samples using a bienzyme electrochemical biosensor with ow injection," Journal of Rapid Methods & Automation in Microbiology, vol. 9, no. 4, pp. 229 { 240, 2001. [45] O. Laczka, J.-M. Maesa, N. Godino, J. del Campo, M. Fougt-Hansen, J. P. Kutter, D. Snakenborg, F.-X. Munoz-Pascual, and E. Baldrich, \Improved bac- teria detection by coupling magneto-immunocapture and amperometry at ow- channel microband electrodes," Biosensors and Bioelectronics, vol. 26, no. 8, pp. 3633 { 3640, 2011. [46] C. Ercole, M. D. Gallo, L. Mosiello, S. Baccella, and A. Lepidi, \Escherichia coli detection in vegetable food by a potentiometric biosensor," Sensors and Actuators B: Chemical, vol. 91, pp. 163 { 168, 2003. [47] A. G. Gehring, D. L. Patterson, and S.-I. Tu, \Use of a light-addressable po- tentiometric sensor for the detection of Escherichia coli O157:H7," Analytical Biochemistry, vol. 258, no. 2, pp. 293 { 298, 1998. [48] G. A. Zelada-Guill en, J. L. Sebasti an-Avila, P. Blondeau, J. Riu, and F. X. Rius, \Label-free detection of Staphylococcus aureus in skin using real-time potentio- metric biosensors based on carbon nanotubes and aptamers," Biosensors and Bioelectronics, vol. 31, no. 1, pp. 226 { 232, 2012. [49] M. dos Santos, J. Agusil, B. Prieto-Simn, C. Sporer, V. Teixeira, and J. Sami- tier, \Highly sensitive detection of pathogen E. coli O157:H7 by electrochemi- cal impedance spectroscopy," Biosensors and Bioelectronics, In Press, Accepted Manuscript, 2013. [50] Y. Luo, S. Nartker, H. Miller, D. Hochhalter, M. Wiederoder, S. Wiederoder, E. Setterington, L. T. Drzal, and E. C. Alocilja, \Surface functionalization of electrospun nano bers for detecting E. coli O157:H7 and BVDV cells in a direct- charge transfer biosensor," Biosensors and Bioelectronics, vol. 26, no. 4, pp. 1612 { 1617, 2010. [51] R. Hao, D. Wang, X. Zhang, G. Zuo, H. Wei, R. Yang, Z. Zhang, Z. Cheng, Y. Guo, Z. Cui, and Y. Zhou, \Rapid detection of Bacillus anthracis using mon- oclonal antibody functionalized QCM sensor," Biosensors and Bioelectronics, vol. 24, no. 5, pp. 1330 { 1335, 2009. [52] D. W. Branch and S. M. Brozik, \Low-level detection of a Bacillus anthracis simulant using Love-wave biosensors on 36 YX LiTaO3," Biosensors and Bio- electronics, vol. 19, no. 8, pp. 849 { 859, 2004. [53] A. P. Davila, J. Jang, A. K. Gupta, T. Walter, A. Aronson, and R. Bashir, \Microresonator mass sensors for detection of Bacillus anthracis Sterne spores in air and water," Biosensors and Bioelectronics, vol. 22, no. 12, pp. 3028 { 3035, 2007. 35 [54] G. A. Campbell and R. Mutharasan, \Detection of Bacillus anthracis spores and a model protein using PEMC sensors in a ow cell at 1 mL/min," Biosensors and Bioelectronics, vol. 22, no. 1, pp. 78 { 85, 2006. [55] L. Fu, K. Zhang, S. Li, Y. Wang, T.-S. Huang, A. Zhang, and Z.-Y. Cheng, \In situ real-time detection of E. coli in water using antibody-coated magne- tostrictive microcantilever," Sensors and Actuators B: Chemical, vol. 150, no. 1, pp. 220 { 225, 2010. [56] J. Wan, M. L. Johnson, R. Guntupalli, V. A. Petrenko, and B. A. Chin, \De- tection of Bacillus anthracis spores in liquid using phage-based magnetoelastic micro-resonators," Sensors and Actuators B: Chemical, vol. 127, no. 2, pp. 559 { 566, 2007. [57] M. Jantunen, O. H anninen, K. Koistinen, and J. Hashim, \Fine PM measure- ments: personal and indoor air monitoring," Chemosphere, vol. 49, no. 9, pp. 993 { 1007, 2002. [58] P. M. Irving, \Air sampling technology: A requirement for biodetection," in Materials of the Second International Symposium on Detection Technologies, Knowledge Foundation, (Arlington, VA, USA), December 2002. [59] http://www.sceptorindustries.com. [60] M. Allan and G. M. Richardson, \Probability density functions describing 24 h inhalation rates for use in human health risk assessments," Human and Ecological Risk Assessment, vol. 4, no. 2, pp. 379 { 408, 1998. [61] C. J. Peters and D. M. Hartley, \Anthrax inhalation and lethal human infection," The Lancet, vol. 359, pp. 710 { 711, 2002. 36 Chapter 3 Phage-Based Magnetoelastic (ME) Biosensors Biosensors can be potential alternatives to the conventional bacterial detection methods. However, as described in the previous chapters, the performance of existing biosensors still needs to be improved. In this chapter, the fundamentals of phage- based ME biosensors, a novel class of wireless, mass-sensitive biosensors, will be described in depth. 3.1 Landscape phages as biomolecular-recognition elements For the past decades, antibodies have been the most commonly used biomolecular- recognition elements [2]. However, their use in on-site bacterial detection might be restricted by such factors as thermal stability, selectivity, and production cost [1,3,4]. Hence, as emerging alternatives, landscape phages have attracted growing attention, and their application to various biosensing systems has been recently reported [3,5{8]. Table 3.1 compares the longevity of a landscape phage and monoclonal antibody spe- ci c for -galactosidase tested at various temperatures [1]. As can be seen, both Table 3.1: Longevity of a landscape phage and monoclonal antibody at various tem- peratures. Temperature Landscape phage Antibody Ref. Room temp. > 6 months > 6 months [1] 37 C 950 days (half-life) 107 days (half-life) [1] 50 C 5 weeks (half-life) 5 weeks [1] 63 C 6 weeks 24 hr [1] 76 C 2.4 days No binding activity [1] 37 phage and antibody retain their binding activities at room temperature for greater than 6 months. However, the antibody degrades much faster than the phage at higher temperatures. For on-site bacterial detection, where a wide range of temperatures are anticipated, a high thermal stability with a fair life time is essential for a biomolecular- recognition element. In addition, as shown in Table 3.2, a high selectivity and low production cost are distinct advantages of landscape phages over antibodies. Table 3.2: Comparison between landscape phages and antibodies in terms of selec- tivity and production cost. Recognition element Selectivity Production cost Ref. Landscape phage High Low [1,3,4] Monoclonal antibody High Very high [1,4] Polyclonal antibody Low High [1,4] Landscape phages are genetically engineered phages that can be synthesized through the phage display technology [9, 10]. This technology, primarily developed Figure 3.1: Schematic illustration of the wild-type fd phage and its genetically engi- neered form, displaying a foreign peptide on the major coat protein pVIII. 38 for the Ff class of lamentous phage strains (i.e., fd, f1, and M13), enables one to construct billions of phage clones that display engineered sequences of peptides on their outer surfaces. For example, Fig. 3.1 shows a schematic illustration of the wild-type fd phage (top) and its genetically engineered form (bottom), displaying a foreign peptide on the major coat protein pVIII. The wild-type Ff phage strains, which possess virtually identical DNA sequences, are exible, thread-like particles about 800 to 900 nm long and 6.5 nm in diameter [1]. They consist of a circular single-stranded DNA ( 6,400 nucleotides) enclosed in a tube of helically arranged molecules of coat proteins (the N-termini exposed on the outer surface and C-termini in the lumen) [10, 11]. There are approximately 2,700 copies of the major coat protein pVIII along the tube?s length, accounting for 98% by mass [1]. In addition, ve copies each of the minor coat proteins cap both ends Figure 3.2: Sequences of amino acid residues of the fd coat proteins. The N-terminus is to the left. The hydrophobic domains are underlined, whereas charged residues are indicated by + or -. 39 of the tube (The minor coat protein pIII and pVI are at one end, whereas the pVII and pIX proteins are at the other end.). Figure 3.2 shows the sequences of amino acid residues of the fd coat proteins (The N-terminus is to the left.) [12]. The sizes of these proteins are: pVIII, 50 residues; pIII, 406 residues; pVI, 112 residues; PVII, 33 residues; and pIV, 32 residues. The N-terminal portion of the protein pIII can be considered as three distinct domains, separated by striking glycine-rich tandem repeat linkers of GGGS and EGGGS between the domains [12]. The numbers within the circles represent amino acid residues assigned to each domain. In the above gure, apolar, hydrophobic domains are underlined, whereas charged residues are indicated by + or -. The major coat protein pVIII, for example, has a hydrophobic domain of continuous 19 amino acid residues (YIGYAWAMVVVIVGATIGI) in the interior of its sequence. Adjacent copies of this protein in the phage virion are held together by hydrophobic interactions between these domains [12]. In addition, the positively charged residues near the C-terminus neutralize the negative charge of the DNA core. Furthermore, all the four minor coat proteins also possess hydrophobic domains similar in length to the hydrophobic domain of pVIII, suggesting that these minor coat proteins may associate with pVIII by hydrophobic interactions [12]. According to Endemann and Model [13], the minor coat proteins pIII, pVI, and pVII all interact with the major coat protein pVIII in phage. Also, the pIII and pIX proteins are exposed to the environment, whereas the pVI and PVII proteins are shielded from the environment. These wild-type Ff phages are viruses that infect the bacterium Escherichia coli bearing F pili. Infection is initiated by the attachment of the N-terminal domain of the pIII protein to the tip of the pilus [10]. As the process continues, the coat proteins dissolve into the surface envelope of the cell, and the viral DNA alone enters the cytoplasm, where a vast number of progeny viral DNA molecules are synthesized by host machinery. These progeny viral DNA molecules are, then, extruded through 40 the cell envelope, acquiring the coat proteins from the cell membrane and emerging as completed virions [10]. Up to 1,000 progeny virions per cell per division can be secreted continuously without killing the host cell [1], leading to a low production cost. The yield of virions can exceed 0.3 mg/mL [10]. In phage display constructions, a foreign coding sequence is spliced in-frame into one of the ve coat protein genes. The resultant foreign peptide encoded by this sequence can be fused to the coat protein and, thereby, displayed on the surface of the virion. In addition, the subsequent length of the phage capsid (i.e., the protein shell of the phage) is altered to match the size of the enclosed recombinant DNA by adding proportionally more pVIII subunits during phage assembly [14]. In this research, three phage clones (E2, JRB7, and SAE10), displaying foreign octamers or nanomers in approximately 4,000 copies of the major coat protein pVIII, were derived from the landscape phage libraries f8/8 and f8/9 [14{16] and used as biomolecular-recognition elements. As shown in Fig. 3.3, three amino acid residues (EGD) of the wild-type pVIII are replaced by a random octamer in the f8/8 library, whereas four residues (EGDD) are replaced by a random nanomer in the f8/9 library, bringing the total size of both fusion pVIII proteins to 55 amino acids. Here, the symbol x represents any amino acid residue. About a half of the pVIII peptide Figure 3.3: Sequences of amino acid residues of the wild-type and fusion pVIII pro- teins. 41 sequence is exposed to the environment, whereas the other half is buried in the capsid [1]. The foreign peptide sequence for each phage clone is shown in Table 3.3. Table 3.3: Phage clones used in this research. Phage Foreign peptide sequence Target pathogen or analyte Ref. E2 VTPPTQHQ S. Typhimurium [15] JRB7 EPRLSPHS B. anthracis [16] SAE10 VPVGAYSDT Streptavidin [14] 3.2 Magnetoelasticity Any magnetic materials exhibit magnetoelastic (ME) behaviour. In other words, the dimensions and elastic properties of these materials are dependent upon their magnetic states, and their magnetic properties are, by contrast, in uenced by internal as well as applied mechanical stresses [17]. Magnetoelasticity has been observed not only in ferromagnets but also in ferrimagnets, antiferrimagnets, paramagnets, and even daimagnets with low susceptibilities [17]. However, from a technical point of view, ferromagnets have been extensively studied for the past centuries. In this work, two amorphous ferromagnets, Metglas Alloy 2826MB and Fe79B21, were used for the construction of ME signal transducers. 3.2.1 Joule magnetostriction An ME material undergoes a change in its dimensions during the process of magnetization. This phenomenon is known as Joule magnetostriction, discovered by and named after James P. Joule [17]. A spherical ME material, for example, may be transformed into an ellipsoid when subjected to an externally applied magnetic eld, H, as illustrated in Fig. 3.4. The induced strains measured in the directions parallel and perpendicular to the eld, k and ?, are largely dependent on the eld strength 42 Figure 3.4: Joule magnetostriction of a spherical ME material. and can be positive or negative, depending on whether the material?s deformation is expansive or compressive. These strains reach their limiting values when the material becomes magnetically saturated. From the hypothetical { H relationships shown in Fig. 3.5, the saturation Joule magnetostriction, s, can be de ned as [18] s = 23( k ?); (3.1) where k and ? are determined by the extrapolation of the tangents of the linear eld dependencies of the strains at the saturated state down to H = 0. s is equal to k when the demagnetized state of the material is isotropic (i.e., k/2 ? = 1). However, for actual materials, the value of k/2 ? is often not equal to 1, depending on both intrinsic parameters (e.g., magnetocrystalline anisotropy) and sample pa- rameters (e.g., demagnetizing eld and internal stresses) [17]. Hence, Eq. 3.1, which 43 remains independent of the demagnetized state of the material, is preferably used to determine s. Figure 3.5: Hypothetical eld dependencies of k and ?. 3.2.2 Magnetization and Joule magnetostriction in ferromagnets When demagnetized at temperatures lower than its Curie temperature, a fer- romagnetic material is divided into a number of magnetic domains to minimize the material?s internal energy as schematically illustrated in Fig. 3.6. Each magnetic do- main is magnetized in a di erent direction such that the net magnetization is zero or small. Between any two adjacent domains, the elementary magnetic moments rotate gradually from one easy magnetization direction (Di) to another (Dj) [17]. When a magnetic eld, H, is applied in any given direction, the following processes occur within the material: 1. The domain D1, whose magnetization direction is the closest to the eld direc- tion, expands at the expense of the domain D3 through the displacement of the 180 domain wall. 44 2. The magnetization in the domain D4 is reversed so that the net magnetization is minimized. 3. When a stronger eld is applied, the domain D1 further expands at the expense of the domains D2 and D4 (90 domain wall displacement), resulting in the formation of only one single domain D1. 4. Finally, the magnetization in this single domain rotates out of its easy magne- tization direction (D1) and aligns along the eld direction. Figure 3.6: Magnetic domains and magnetization processes in a ferromagnet. During the above magnetization processes, the material is strained due to the ME coupling since the distribution of the elementary magnetic moments becomes anisotropic (upper left illustration in Fig. 3.7a). This anisotropic distribution of mag- netic moments may also be induced when an ME material is mechanically stressed (upper right illustration in Fig. 3.7a). As a result, spontaneous Joule magnetostric- tion is also induced, again due to the ME coupling. When a magnetic eld is, then, applied to this material, k becomes a function of the applied pre-stress [17]. In the case of positive Joule magnetostriction, a compressive stress orients the mag- netic moments into the direction perpendicular to the stress direction. Such eld 45 and stress dependence of k is shown in Fig. 3.7b. As can be seen, k is increased with larger pre-stresses, which rotate more magnetic moments (Note that s is, by contrast, usually nearly stress-independent [17].). Figure 3.7: (a) E ects of magnetizing eld and mechanical stress on the distribu- tion of magnetic moments in a ferromagnet and (b) eld dependence of k under a compressive stress. Figure 3.8: Temperature dependence of normalized s in Fe80B20. 46 Joule magnetostriction is also temperature dependent. The thermal dependence of s is usually monotonous as in the case for Fe80B20, shown in Fig. 3.8, where Tc is the Curie temperature [17]. 3.2.3 ME signal transducers An ideal material for the construction of ME signal transducers is a magnetically soft material that possesses a high saturation Joule magnetostriction, s, and a high magnetomechanical coupling factor, k, which can be de ned as [17] k = EmepE eEm ; (3.2) whereEme, Ee, andEm are the mutual elastic and magnetic energy density, elastic self- energy density, and magnetic self-energy density, respectively. This coupling factor can be used to characterize the material?s ability to convert a magnetic energy into an elastic energy and vice versa. Traditionally, Metglas Alloy 2826MB (from Honeywell International) with a composition of Fe40Ni38Mo4B18 has been the material for ME signal transducers. This amorphous, ferromagnetic alloy is mechanically robust (e.g., tensile strength of 1 to 2 GPa), and it possesses reasonably high s and k values [19{22]. In addition, the alloy has a low material cost, allowing ME signal transducers made of this alloy to be used on a disposable basis [23]. In recent years, another amorphous, ferromagnetic alloy with a composition close to Fe80B20 has also been used for the construction of ME signal transducers [7, 24]. This iron-rich alloy can be easily produced through physical or electrochemical de- position processes. Hence, by combining with standard microelectronic fabrication techniques, batch fabrication of miniature ME signal transducers is also possible, which further reduces the fabrication cost. Some important materials properties for the above alloys are summarized in Table 3.4. 47 Table 3.4: Materials properties for Metglas 2826MB and Fe80B20. Property Metglas 2826MB Ref. Fe80B20 Reference Elastic modulus [GPa] 100 to 110 [22] 166 [25] Density [ 103 kg/m3] 7.9 [22] 7.4 [25] Poisson?s ratio 0.33 [26] 0.3 [27] Saturation magnetostriction, s [ppm] 12 [22] 32 [28] Magnetomechanical coupling fac- tor, k 0.98 [23] 0.64 [29] 3.3 Fabrication of ME sensor platforms Depending on the size of sensor platforms (i.e., ME signal transducers), two dif- ferent fabrication methods were employed. A dicing method was used for millimeter- scale sensor platforms (0.5 to 4-mm long), while a co-sputtering-based method was used for micron-scale sensor platforms (100 to 500- m long). 3.3.1 Dicing method A ribbon of Metglas Alloy 2826MB was purchased from Honeywell International. Small pieces with a size of 50 mm 12.7 mm 30 m were cut from the ribbon and double-side polished down to a thickness of 15 m. The polished pieces were, then, diced into millimeter-scale strip-shaped sensor platforms (Fig. 3.9) using a automated dicing saw. After cleaning with acetone and ethanol in an ultrasonic bath, these diced sensor platforms were successively coated with thin layers of Cr (90 nm) and Au (150 nm) by electron-beam induced deposition. The Cr layer acts as an adhesive interlayer between the Au layer and sensor platform. The Au layer provides corrosion resistance as well as a ready surface for the immobilization of the phages. 48 Figure 3.9: Diced sensor platforms stored in dry methanol. 3.3.2 Co-sputtering-based method Micron-scale ME sensor platforms were batch-fabricated using a co-sputtering- based method reported previously [7]. The fabrication procedure is diagrammed in Fig. 3.10. First, a four-inch gold-coated wafer was photolithographically patterned with 10- m thick rectangular islands of the STR-1045 photoresist (from Rohm and Haas Electronic Materials, LLC). The lateral dimensions of these photoresist islands were kept the same as those of sensor platforms of target size. Next, the patterned wafer was successively deposited with 50-nm thick Au, 4- m thick Fe79B21 (i.e., co- sputtering of Fe and B), and another 50-nm thick Au using a Denton sputter coater (Fig. 3.11). Then, the Au-enclosed Fe79B21 alloy on the photoresist islands was lifted o the wafer to become freestanding ME sensor platforms by dissolving the underlying photoresist with acetone. This method not only ensures the dimensional 49 Figure 3.10: Procedure for the co-sputtering-based method. consistency of the fabricated sensor platforms but also greatly reduces the fabrication cost per sensor platform due to the batch fabrication process (See Fig. 3.12, showing batch-fabricated sensor platforms with a size of 100 m 25 m 4 m.). 50 Figure 3.11: Denton sputter coater. 51 Figure 3.12: Scanning electron micrograph of batch-fabricated sensor platforms with a size of 100 m 25 m 4 m on a gold-coated wafer. The sputtering conditions used are summarized in Table 3.5. Prior to sputtering, the chamber was pumped down to 7 10 7 Torr in order to minimize residual oxygen in the fabricated sensor platforms. In addition, during the deposition process, the sample stage was rotated so that the uniformity of the deposits can be guaranteed. Table 3.5: Sputtering conditions used for the fabrication of micron-scale sensor plat- forms. Target Cathode type Power [W] Time [sec] Ar ow rate [sccm] Cr DC 50 100 25 Au DC 100 200 3 times 12 Fe DC 41 64,000 30 B RF 100 64,000 30 52 3.3.3 Annealing The fabricated sensor platforms were nally annealed in vacuum at 220 C for 2 h to relieve residual internal stresses and minimize the e ects of any surface defects from the fabrication processes [30]. This temperature was chosen because it is high enough to e ectively anneal the sensor platforms within a reasonable time, yet it is still far below the Curie temperatures as well as the recrystallization temperatures of the ME materials [7]. 3.3.4 Fabrication cost per sensor platform As discussed in Chapter 1, cost is an important factor for the on-site detection of pathogens because proper allocation of limited resources is essential to improving overall safety. Through the batch fabrication of micron-scale sensor platforms, great reduction in fabrication cost can be realized. For a rough estimation of the fabrication cost per sensor platform, the following assumptions were made: 1. The size of sensor platforms to be fabricated is 200 m 40 m 4 m, and their composition is Fe79B21. 2. The size of a silicon wafer on which sensor platforms will be fabricated is four inches in diameter. 80% of the wafer surface area will be used. 3. Only the material costs for metals to be sputtered (i.e., Cr, Au, Fe, and B as shown in Table 3.6) and the silicon wafer ( a3620) are considered. In other words, the costs for the STR-1045 photoresist and other required chemicals are not considered. 4. The metals will be sputter-deposited only on the wafer. 5. The electricity costs for photolithography, annealing, and other fabrication pro- cesses are not considered. Only the energy charges for sputtering are considered. 53 An electricity rate of 7.49a162/kWh paid by Auburn University in FY 2009 [31] is used. Table 3.6: Material costs for the sputtering targets. Material Cost [a36/g] Density [g/cm3] Deposit?s thickness Ref. Cr 1 7.2 10 nm [32] Au 75 19.3 150 nm (50 nm 3) [33] Fe 2 7.9 3.2 m (79% of 4 m) [34] B 2 2.3 0.8 m (21% of 4 m) [35] Hence, the fabrication cost per sensor platform, Csensor, can be calculated by Csensor = material costs + energy chargesnumber of sensors = X (Mi i Vi) + wafer cost + X (Pi Ti $0:0749=kWh) =Nsensor; (3.3) i = Cr; Au; Fe; or B; where M, , and V represent the material cost (per gram), density, and volume for each sputtered metal; P and T denote the power (in kilo-watts) and time (in hours) for sputtering; and Nsensor is the number of sensors to be fabricated on the wafer. By using the values shown in Tables 3.5 and 3.6, Csensor = ($5:8 10 4 + $1:8 + $4:1 10 1 + $3:0 10 2 + $20 + $1:0 10 4 + $1:3 10 3 + $5:5 10 2 + $1:3 10 1)=8:1 105 sensors $2:8 10 5: (3.4) 54 3.4 Fabrication of phage-based ME biosensors 3.4.1 Immobilization of a phage on the ME sensor platforms The annealed sensor platforms were individually immersed in 330 l of a phage suspension (usually, phage in a tris-bu ered saline (TBS) bu er solution, 5 1011 vir/ml) in a polypropylene PCR tube. The tubes were, then, rotated with a Barn- stead LabQuake tube rotator (from Fisher Scienti c, Inc.) at 8 rpm for 1 h. In this way, the phage was allowed to uniformly attach to platform surfaces via phys- ical adsorption. Finally, these phage-immobilized ME biosensors (i.e., measurement sensors) were thoroughly rinsed with sterile distilled water to remove any TBS bu er components as well as loosely attached phages from the platform surfaces. Covalent immobilization of a phage on the surface of sensor platforms will be described in Chapter 6. 3.4.2 Surface blocking of the ME biosensors with bovine serum albumin In order to reduce non-speci c adsorption of S. Typhimurium cells or B. anthracis spores on biosensor surfaces, surface blocking with bovine serum albumin (BSA) was performed. The prepared measurement sensors were individually immersed in a 330- l solution of BSA (0.01 to 1 % w/v in sterile distilled water) in a PCR tube. After 40 min of tube rotation at 8 rpm, the biosensors were collected from the solution and thoroughly rinsed with sterile distilled water to be ready for use. Control sensors, which are not immobilized with phage but only surface-blocked with BSA, were also prepared and used for background subtraction. 55 3.5 Principle of detection The fabricated ME biosensors are made from one of the magnetostrictive alloys (i.e., Metglas 2826MB or Fe79B21). Hence, the biosensors can be placed into magneto- mechanical resonance when subjected to an externally applied magnetic eld that alternates at the right frequency. For a freestanding, strip-shaped biosensor, the fundamental resonant frequency of longitudinal vibration, f, can be expressed by [26] f = 12L s E (1 ); (3.5) where L, E, , and denote the length, modulus of elasticity, density, and Poisson?s ratio of the biosensor, respectively. When the biosensor and bacterial cells (or spores) come into contact with each other, the phage that is immobilized on the biosensor binds with the cells (or spores) (Fig. 3.13), thereby increasing the total mass of the biosensor. This change in mass causes a corresponding decrease in the biosensor?s resonant frequency, which is the principle of detection. In addition, for uniform mass attachment, the mass sensitivity of the biosensor, Sm, de ned as the ratio of the resonant frequency change, f, to the mass change m, can be approximated Figure 3.13: Uniform attachment of bacterial cells or spores on a phage-immobilized ME biosensor. 56 by [23,36] Sm = f m 14L2WT s E 3(1 ); (3.6) where W and T represent the width and thickness of the biosensor, respectively. Equation 3.6 describes that the mass sensitivity is largely dependent on the size of the biosensor and inversely proportional to L2WT. 3.5.1 Minimum detectable number of bacterial cells Table 3.7 shows ME biosensors of di erent size and their theoretical detection limits in terms of the minimum detectable number of bacterial cells, Nmin. The rst four biosensors from the top are made of Metglas 2826MB, whereas the rest is made of Fe79B21. By rearranging Eq. 3.6, the minimum detectable mass change, mmin, can be calculated by mmin = fminS m ; (3.7) where fmin represents the corresponding minimum detectable frequency change. If the mass of a single bacterial cell is assumed to be 1 pg [37], and a frequency shift of 1,000 Hz (i.e., fmin = 1,000 Hz) can be resolved, the minimum detectable number of bacterial cells, Nmin, can then be computed by Nmin = mmin1 pg = fmin Sm 1 pg = 1;000 Hz 1 pg 1 Sm; (3.8) where Nmin is an integer. As can be seen in the table, detection of a single bacterial cell may be possible with a 50 m 10 m 2 m sensor. 57 Table 3.7: Di erently sized ME biosensors and their theoretical detection limits. L [ m] W [ m] T [ m] Sm [Hz/pg] mmin [pg] Nmin Material 4,000 800 15 7.34 10 4 1.36 106 1,362,212 Metglas 2,000 400 15 5.87 10 3 1.70 105 170,277 Metglas 1,000 200 15 4.70 10 2 2.13 104 21,285 Metglas 500 100 15 3.76 10 1 2.66 103 2,661 Metglas 500 250 4 7.65 10 1 1.31 103 1,308 Fe79B21 500 167 4 1.15 8.73 102 874 Fe79B21 500 125 4 1.53 6.54 102 654 Fe79B21 500 100 4 1.91 5.23 102 523 Fe79B21 200 100 4 1.20 101 8.37 101 84 Fe79B21 200 67 4 1.78 101 5.61 101 57 Fe79B21 200 50 4 2.39 101 4.18 101 42 Fe79B21 200 40 4 2.99 101 3.35 101 34 Fe79B21 150 75 4 2.83 101 3.53 101 36 Fe79B21 150 50 4 4.25 101 2.35 101 24 Fe79B21 150 38 4 5.59 101 1.79 101 18 Fe79B21 150 30 4 7.08 101 1.41 101 15 Fe79B21 100 50 4 9.56 101 1.05 101 11 Fe79B21 100 33 4 1.45 102 6.90 7 Fe79B21 100 25 4 1.91 102 5.23 6 Fe79B21 100 20 4 2.39 102 4.18 5 Fe79B21 50 10 2 3.82 103 0.26 1 Fe79B21 58 3.5.2 Measurement of the resonant frequency of the ME biosensors In order to determine the resonant frequency of the biosensors, they were in- dividually placed into the center of a copper solenoid coil that is connected to a network analyzer (HP/Agilent 8751A from Agilent Technologies, Inc.), operated in the S11 re ection mode (Fig. 3.14). An incident AC signal was, then, applied across the coil to magnetically excite the longitudinal vibration of the biosensor, and the resultant re ected signal was compared with the incident signal over a proper range of frequencies. Finally, the resonant frequency of the biosensor was determined to be the frequency at which the largest change in normalized jS11j occurs, due to the magneto-mechanical resonance of the biosensor. To enhance the magnitude of the resonance peak, a proper bias magnetic eld was also applied to the biosensor with a bar magnet. Figure 3.15 shows a response of a typical 500 m 100 m 4 m ME biosensor in air. Figure 3.14: Measurement setup. 59 Figure 3.15: Response of a typical 500 m 100 m 4 m ME biosensor in air. 60 Bibliography [1] V. A. Petrenko, \Landscape phage as a molecular recognition interface for de- tection devices," Microelectronics Journal, vol. 39, no. 2, pp. 202 { 207, 2008. [2] S. Goodchild, T. Love, N. Hopkins, and C. Mayers, \Engineering antibodies for biosensor technologies," vol. 58 of Advances in Applied Microbiology, pp. 185 { 226, Academic Press, 2005. [3] J. Wan, M. L. Johnson, R. Guntupalli, V. A. Petrenko, and B. A. Chin, \De- tection of Bacillus anthracis spores in liquid using phage-based magnetoelastic micro-resonators," Sensors and Actuators B: Chemical, vol. 127, no. 2, pp. 559 { 566, 2007. [4] V. A. Petrenko and I. B. Sorokulova, \Detection of biological threats. a challenge for directed molecular evolution," Journal of Microbiological Methods, vol. 58, pp. 147 { 168, 2004. [5] C. Mao, A. Liu, and B. Cao, \Virus-based chemical and biological sensing," Angewandte Chemie International Edition, vol. 48, pp. 6790 { 6810, 2009. [6] J. Wan, H. Shu, S. Huang, B. Fiebor, I.-H. Chen, V. A. Petrenko, and B. A. Chin, \Phage-based magnetoelastic wireless biosensors for detecting Bacillus anthracis spores," IEEE Sensors Journal, vol. 7, pp. 470 { 477, 2007. [7] M. L. Johnson, J. Wan, S. Huang, Z. Cheng, V. A. Petrenko, D.-J. Kim, I.- H. Chen, J. M. Barbaree, J. W. Hong, and B. A. Chin, \A wireless biosensor using microfabricated phage-interfaced magnetoelastic particles," Sensors and Actuators A: Physical, vol. 144, no. 1, pp. 38 { 47, 2008. [8] R. S. Lakshmanan, R. Guntupalli, J. Hu, D.-J. Kim, V. A. Petrenko, J. M. Barbaree, and B. A. Chin, \Phage immobilized magnetoelastic sensor for the de- tection of Salmonella typhimurium," Journal of Microbiological Methods, vol. 71, no. 1, pp. 55 { 60, 2007. [9] V. A. Petrenko, G. P. Smith, X. Gong, and T. Quinn, \A library of organic landscapes on lamentous phage," Protein Engneering, vol. 9, pp. 797 { 801, 1996. [10] G. P. Smith and V. A. Petrenko, \Phage display," Chemical Reviews, vol. 97, pp. 391 { 410, 1997. 61 [11] V. Petrenko, G. Smith, M. Mazooji, and T. Quinn, \ -helically constrained phage display library," Protein Engineering, vol. 15, no. 11, pp. 943 { 950, 2002. [12] D. Marvin, \Filamentous phage structure, infection and assembly," Current Opinion in Structural Biology, vol. 8, no. 2, pp. 150 { 158, 1998. [13] H. Endemann and P. Model, \Lcoation of lamentous phage minor coat proteins in phage and in infected cells," Journal of Molecular Biology, vol. 250, no. 4, pp. 496 { 506, 1995. [14] G. Kuzmicheva, P. Jayanna, I. Sorokulova, and V. Petrenko, \Diversity and cen- soring of landscape phage libraries," Protein Engineering, Design and Selection, vol. 22, pp. 9 { 18, 2009. [15] I. B. Sorokulova, E. V. Olsen, I.-H. Chen, B. Fiebor, J. M. Barbaree, V. J. Vodyanoy, B. A. Chin, and V. A. Petrenko, \Landscape phage probes for Salmonella typhimurium," Journal of Microbiological Methods, vol. 63, no. 1, pp. 55 { 72, 2005. [16] J. Brigati, D. D. Williams, I. B. Sorokulova, V. Nanduri, I.-H. Chen, C. L. Turnbough, and V. A. Petrenko, \Diagnostic probes for Bacillus anthracis spores selected from a landscape phage library," Clinical Chemistry, vol. 50, no. 10, pp. 1899 { 1906, 2004. [17] E. du Tr emolet de Lacheisserie, Magnetostriction: Theory and Applications of Magnetoelasticity. CRC Press, 1993. [18] G. Vlas ak, \Direct measurement of magnetostriction of rapidly quenched thin ribbons," Journal of Magnetism and Magnetic Materials, vol. 215-216, pp. 479 { 481, 2000. [19] A. Hernando, M. Vazquez, and J. M. Barandiaran, \Metallic glasses and sensing applications," Journal of Physics E: Scienti c Instruments, vol. 21, pp. 1129 { 1139, 1988. [20] C. Modzelewski, H. T. Savage, L. T. Kabaco , and A. E. Clark, \Magnetome- chanical coupling and permeability in transversely annealed Metglas 2605 alloys," IEEE Transactions on Magnetics, vol. 17, pp. 2837 { 2839, 1981. [21] R. C. O?Handley, Modern Magnetic Materials: Principles and Applications. New York: John Wiley & Sons Inc., 2000. [22] http://www.metglas.com/products/page5 1 2 7.htm. [23] C. A. Grimes, C. S. Mungle, K. Zeng, M. K. Jain, W. R. Dreschel, M. Paulose, and K. G. Ong, \Wireless magnetoelastic resonance sensors: A critical review," Sensors, vol. 2, pp. 294 { 313, 2002. 62 [24] S. Li, Development of novel acoustic wave biosensor platforms based on magne- tostriction and fabrication of magnetostrictive nanowires. PhD thesis, Auburn University, 2007. [25] L. Davis, R. Ray, C.-P. Chou, and R. O?Handley, \Mechanical and thermal properties of Fe80B20 glass," Scripta Metallurgica, vol. 10, pp. 541 { 546, 1976. [26] C. Liang, S. Morshed, and B. C. Prorok, \Correction for longitudinal mode vibration in thin slender beams," Applied Physics Letters, vol. 90, p. 221912, 2007. [27] H.-U. K unzi, \Mechanical properties of metallic glasses," in Glassy Metal II, vol. 53 of Topics in Applied Physics, pp. 169{216, Springer Berlin Heidelberg, 1983. [28] C. China, \Magnetic properties of sputter deposited Fe based amorphous thin lms for resonator application," Master?s thesis, the University of Central Florida, 2006. [29] M. A. Mitchell, A. E. Clark, H. T. Savage, and R. J. Abbundi, \ E e ect and magnetomechanical coupling factor in Fe80B20 and Fe78Si10B12 glassy ribbons," IEEE Transactions on Magnetics, vol. 14, pp. 1169 { 1171, 1978. [30] S. Huang, J. Hu, J. Wan, M. Johnson, H. Shu, and B. Chin, \The e ect of annealing and gold deposition on the performance of magnetoelastic biosensors," Materials Science and Engineering: C, vol. 28, no. 3, pp. 380 { 386, 2008. [31] http://rs.acupcc.org/site media/uploads/cap/699-cap 1.pdf. [32] http://www.lesker.com/newweb/deposition materials/depositionmaterials sputtertargets 1.cfm?pgid=cr1. [33] http://www.lesker.com/newweb/deposition materials/depositionmaterials sputtertargets 1.cfm?pgid=au1. [34] http://www.lesker.com/newweb/deposition materials/depositionmaterials sputtertargets 1.cfm?pgid=fe1. [35] http://www.lesker.com/newweb/deposition materials/depositionmaterials sputtertargets 1.cfm?pgid=bor1. [36] C. A. Grimes, S. C. Roy, S. Rani, and Q. Cai, \Theory, instrumentation and applications of magnetoelastic resonance sensors: A review," Sensors, vol. 11, no. 3, pp. 2809 { 2844, 2011. [37] http://hypertextbook.com/facts/2003/LouisSiu.shtml. 63 Chapter 4 Direct Detection of S. Typhimurium on Fresh Spinach Leaves 4.1 Introduction Frequent outbreaks of bacterial food poisoning as well as associated illnesses are signi cant public health concerns. Hence, to secure the safety of our food supply, proper, routine testing of food products must be implemented throughout the sup- ply chain. Conventionally, such testing has been conducted with phenotypic and/or genotypic assays, including culture- and PCR-based assays [1{10]. Although capable of providing con rmatory results with low detection limits, these methods are often complex, expensive, labour-intensive, and thus, not suitable for the on-site detection of pathogenic bacteria. In addition, rapidness of the testing is among the most impor- tant requirements particularly for dealing with a large volume of fresh produce and other essential food items that are consumed daily. To overcome the drawbacks of the conventional detection methods and facilitate on-site bacterial detection, much re- search has been recently focused on developing label-free biosensors [11{19]. However, even for these biosensors, sample preparation, including the collection, puri cation, and enrichment of a pathogen-containing sample, is generally required prior to the testing. Hence, there is a motivation for eliminating any pre-test sample preparation steps to simplify the procedure and further reduce the total time and cost of testing. Attempts using phage-based ME biosensors have been recently made for the rapid, direct detection of S. Typhimurium on fresh produce (i.e., tomatoes and shell eggs [20{22]). These biosensors are composed of a freestanding, strip-shaped ME signal transducer coated with the E2 phage [23], which speci cally binds with S. Typhimurium. These biosensors can be directly placed on produce surfaces due to 64 their wireless, freestanding nature and used to monitor the presence of the bacterium without pre-test sample preparation. In this investigation, the methodology was employed to test Salmonella-spiked spinach leaves. Figures 4.1a to 4.1c show scanning electron micrographs of the surfaces of various produce: (a) a tomato, (b) a shell egg, and (c) a spinach leaf. As can be seen, surface topography varies from one produce to another. In addition, since spinach leaves are likely to possess complex surface topography as shown in Fig. 4.1d (a close-up view of a leaf surface), these surface features may a ect the physical contact between the Figure 4.1: Scanning electron micrographs of various produce surfaces: (a) tomato, (b) eggshell, and (c) spinach leaf. A close-up view of a spinach leaf spiked with S. Typhimurium is shown in (d). 65 biosensors and S. Typhimurium cells. Hence, three di erent sizes of biosensors (2- mm, 500- m, and 150- m long) were fabricated and tested to investigate the e ects of sensor size on the limit of detection (LOD). Furthermore, a formula describing the probability of detection as a function of the size and number of biosensors and the surface density of S. Typhimurium was derived. By using the formula, the required number of biosensors to obtain a desired LOD can be determined. 4.2 Material and methods 4.2.1 E2 phage and S. Typhimurium Suspensions of the E2 phage (5 1011 virions/ml in a TBS bu er) and S. Ty- phimurium cells (ATCC 13311 at a concentration of 5 108 cells/ml in sterile dis- tilled water) were provided by Dr. James Barbaree?s group at Auburn University. The concentrated Salmonella suspension was diluted with sterile distilled water as desired prior to use. 4.2.2 Confocal re ectance imaging of spinach leaf surfaces Both adaxial and abaxial surfaces of fresh spinach leaves were imaged with a confocal scanning laser microscope (Nikon A1 from Nikon Corp.). The leaves were cut into 10 mm 10 mm pieces without including any major leaf veins and mounted on a glass slide with double-sided tape. These samples were prepared right before the imaging and individually surface-scanned with a 488-nm line laser at 21 C and 37% relative humidity. To minimize the degradation of the samples, a low laser power of 1.5% was used exclusively, and the imaging per sample was completed within 30 min. Re ectance from the sample surfaces was collected at magni cations of 40 and 400 with a bandpass lter of 482/35 nm. The collected digital data were saved as images with a pixel resolution of 1,024 1,024. The unit pixel length was 3.11 m and 0.311 m for images taken at 40 and 400 magni cations, respectively. 66 To reconstruct and characterize the topography of the sample surfaces, a series of plane images through the thickness of spinach leaf surfaces was captured. The separation between adjacent slices was 0.1 m. Fifteen samples were prepared for both adaxial and abaxial surfaces and imaged at the above-mentioned magni cations. The ImageJ software with the SurfCharJ plugin [24] was, then, used to reconstruct the surface topography and extract surface height data associated with their location of pixels. Finally, surface pro les were regenerated and tilt-corrected by subtracting an overall increasing or decreasing linear trend along the sampling length. 4.2.3 Fabrication of ME sensor platforms with three di erent sizes Freestanding, strip-shaped ME sensor platforms with three di erent sizes were fabricated of either Metglas 2826MB or Fe79B21, both of which are amorphous, fer- romagnetic alloys with ME properties [25,26]. Millimeter-long sensor platforms with two di erent sizes (2 mm 0.4 mm 15 m and 0.5 mm 0.1 mm 15 m) were manufactured by polishing and dicing a sheet of Metglas 2826MB. By contrast, Figure 4.2: Di erently sized sensor platforms used (top view). 67 micrometer-long sensor platforms with a size of 150 m 30 m 4 m were fabri- cated of Fe79B21 using the co-sputtering-based method described in Chapter 3. Figure 4.2 illustrates the three di erently sized sensor platforms used in this investigation. 4.2.4 Fabrication of phage-based ME biosensors The fabrication of phage-based ME biosensors was completed by following the procedures described in Chapter 3. In addition to measurement sensors, control sensors, which are not immobilized with the E2 phage but only surface-blocked with BSA, were also prepared and used for background subtraction. 4.2.5 Determination of the concentration of BSA for surface blocking In order to determine a reasonable concentration of BSA for surface blocking, measurement and control sensors (2-mm long) exposed to three di erent concentra- tions of BSA (0.01, 0.1, and 1 % w/v) were prepared and tested. These sensors were placed on a wet spinach leaf surface inoculated with S. Typhimurium (5 108 cells/ml, 40 L) and, then, allowed for binding. Figure 4.3a shows responses for both measurement and control sensors. As can be seen, the resonant frequency changes for both types of sensors decreased with increased concentrations of BSA. Hence, in terms of surface blocking, a high BSA concentration is preferable. However, the re- sponse of control sensors must be minimized without unnecessarily reducing that of measurement sensors. Hence, it is important to nd a BSA concentration at which the di erence between the responses of measurement and control sensors is statisti- cally maximized. For the three pairs of data in Fig. 4.3a, the con dence level of di erence was calculated with a standard t-test [27,28]. The best result (i.e., highest con dence level of di erence) was obtained at a BSA concentration of 0.1 % w/v as shown in Fig. 4.3b. This concentration was, hence, used for surface blocking. 68 Figure 4.3: E ects of BSA concentration on (a) resonant frequency changes for mea- surement and control sensors (2-mm long) and on (b) the con dence level of di erence. 4.2.6 Direct detection of S. Typhimurium on fresh spinach leaves Pre-washed, bagged, fresh baby spinach leaves (Kroger brand) were purchased from a local grocery store and used as-received. They were, rst, individually ad- hered to a clean, at surface with double-sided tape. Forty-microliter drops of S. Typhimurium with various concentrations (i.e., ten-fold serial dilutions of 5 108 cells/ml) were, then, spot-inoculated on the leaf surface as illustrated in Fig. 4.4a. Any major leaf veins, which cause a sudden, large change in surface topography, were avoided. Yet, the total surface area of the major leaf veins was found to be small (7.3 69 Figure 4.4: Schematic illustration of the test procedure: (a) spot-inoculation of S. Typhimurium on the leaf surface and measurement of the initial resonant frequency of biosensors, (b) placement of both measurement and control sensors on the Salmonella- inoculated sites (after drying the Salmonella drops and misting the leaf surface), (c) measurement of the nal resonant frequency of the biosensors, and (d) typical responses of the biosensors. 2.9 % of a leaf). The inoculated Salmonella drops were, then, allowed to dry in air for 90 to 120 min. Next, the leaf surface was uniformly misted by spraying ster- ile distilled water ( 20 l/cm2), and both measurement and control sensors with a pre-determined resonant frequency were directly placed on the Salmonella-inoculated sites (Fig. 4.4b). After 25 min to allow for binding, the biosensors were collected with a magnet, and measurement of their nal resonant frequency was completed within 20 min (Figs. 4.4c and 4.4d) (The method for resonant frequency measurement was described in Chapter 3). The total test time was, hence, roughly 45 min. In this investigation, the test was performed at 23 C and 35% relative humidity. The three 70 di erently sized biosensors (i.e., 2 mm-, 0.5 mm-, and 150 m-long sensors) were used to test both adaxial and abaxial surfaces of the leaves. Ten measurement and control sensors each were used for each concentration of S. Typhimurium. In order to convert cells/ml into cells/cm2 (i.e., surface density of S. Typhimurium), the area of the inoculation sites was measured and found to be 0.22 0.02 cm2 and 0.30 0.03 cm2 for the adaxial and abaxial surfaces, respectively. 4.3 Results 4.3.1 Observation of Salmonella-inoculated leaf surfaces Figure 4.5: Scanning electron micrographs of a spinach leaf surface inoculated with a 40- l drop of S. Typhimurium with various concentrations: (a) 5 108 cells/ml, (b) 5 107 cells/ml, (c) 5 106 cells/ml (with a 150 m-long ME biosensor), and (d) 0 cells/ml (reference). 71 Figure 4.5 shows representative scanning electron micrographs of a spinach leaf surface inoculated with a 40- l drop of S. Typhimurium with various concentrations. At a concentration of 5 108 cells/ml, the leaf surface was nearly completely covered by S. Typhimurium cells (Fig. 4.5a). The number of observable cells, then, decreased with decreased concentrations of inoculated cells as anticipated, and the distribution of cells became non-uniform for lower concentrations (Figs. 4.5b and 4.5c). This localization of cells may be attributed to localized availability of nutrients as well as leaf surface conditions (e.g., hydrophobicity and surface charges), which a ect the motility and attachment of the cells [29,30]. Furthermore, the leaf surface was found to possess complex topography (Fig. 4.5d), which plays a crucial role in physical contact between a biosensor and S. Typhimurium cells. 4.3.2 Resonant frequency measurement Figure 4.6 shows a response of a typical 150 m 30 m 4 m sensor, which was measured in air with the setup described in Chapter 3. It can be seen that the raw data set (dim gray curve in Fig. 4.6a) contains a high degree of noise even after a 10-time averaging operation. This noisy raw curve is common particularly for small sensors with small peak amplitudes, which requires additional smoothing of the data set. Hence, to further reduce the noise level, the Savitzky-Golay smoothing (25 points, second-order) was performed (black curve in Fig. 4.6a). In addition, Lorentzian tting of the smoothed curve was nally performed to determine the resonant frequency of the sensor as shown in Fig. 4.6b. This method of data arrangement was also used for larger sensors. The mean resonant frequencies for the three di erently sized biosensors are summarized in Table 4.1. 72 Figure 4.6: Response of a typical 150 m 30 m 4 m sensor in air: (a) raw data set (10-time averaged) and its smoothed curve and (b) Lorentizan tting of the smoothed curve. 73 Table 4.1: Mean resonant frequencies for the di erently sized biosensors. Sensor size Mean resonant frequency 2 mm 0.4 mm 15 m 1.12 0.04 MHz 0.5 mm 0.1 mm 15 m 4.41 0.04 MHz 150 m 30 m 4 m 13.06 0.12 MHz To determine a change in the resonant frequency of a sensor, both the initial and nal resonant frequencies (finitial and f nal) need to be measured. An example is shown in Fig. 4.7, where the resonant peaks for a 150- m long sensor before and after placing on a leaf surface inoculated with S. Typhimurium (5 108 cells/ml) are shown. It can be seen that the nal peak appears at a lower frequency due to the attachment of S. Typhimurium cells. Figure 4.7: Resonant peaks for a 150- m long sensor before and after placing on a leaf surface inoculated with S. Typhimurium at a concentration of 5 108 cells/ml. 4.3.3 Dose-response of the ME biosensors Dose-response relationships for the di erently sized biosensors are shown in Fig. 4.8. The plots on the left and right are the results for the adaxial and abaxial surfaces 74 Figure 4.8: Dose-response plots for the di erently sized biosensors (2 mm-, 0.5 mm- and, 150 m-long sensors). The plots on the left and right are the results for the adaxial and abaxial surfaces, respectively. of spinach leaves, respectively. Resonant frequency changes of measurement sensors (circles) were found to be largely dependent on the surface density of S. Typhimurium. By contrast, control sensors (squares) showed much smaller responses, indicating that selective binding of S. Typhimurium on the measurement sensors occurred. In 75 addition, the standard error was found to be large at high surface densities of S. Typhimurium, which may be attributed to (1) the complex topography of leaf surfaces and (2) the random locations of the sensors, resulting in non-uniform, inconsistent physical contact between the sensors and S. Typhimurium cells. 4.3.4 Determination of the LOD The LODs for the above dose-response plots were determined as follows: 1. The responses of the control sensors were subtracted from those of the measure- ment sensors (i.e., background subtraction). Figure 4.9: (a) Sigmoidal curve and (b) the determination of the LOD. 76 2. The resultant data were curve- tted with sigmoidal functions, which are com- monly used for the description of the response patterns of bioassays [31]. On a sigmoidal curve, there are usually two concentration-independent regions (i.e., lower and upper plateaus) and one linear concentration-dependent region de- ned with two bend points as shown in Fig. 4.9a. 3. Finally, the LOD was determined as the concentration at which the tted re- sponse deviates from the average response in the lower plateau region, fAVE, by a multiple of the standard error, , in the region [32]. The value of the multiple is dependent on a required statistical signi cance level. In this work, a multiple of three was used (i.e., 3 ) as shown in Fig. 4.9b. Note that the lower plateau region was determined through a linear regression analysis with an R2 value of greater than 0.95. Figure 4.10 shows background-subtracted data for the dose-response plots shown in Fig. 4.8. These data were tted with sigmoidal functions (red solid curves), which can be de ned as [31] Y = a b1 + (X=c)d +b; (4.1) where Y, X, a, b, c, and d represent the response, concentration (i.e., surface density of S. Typhimurium), lower asymptote, upper asymptote, in ection point, and slope factor, respectively. The R2 values were found to be all close to the unity, indicating that the curve tting was well performed. To determine the LOD, the fAVE + 3 values were extrapolated to the tted curves as shown in Fig. 4.10 (blue dashed lines). Table 4.2 summarizes the LODs of the di erently sized biosensors for both adaxial and abaxial surfaces of spinach leaves. 77 Figure 4.10: Background-subtracted data for the dose-response plots in Fig. 4.8. These data were tted with sigmoidal functions (red solid curves). The R2 values were all close to one. The values of fAVE + 3 are shown in blue text. As can be seen in Table 4.2, the LOD was found to be on the order of 104 to 105 cells/cm2. The best results were obtained with the 150- m long sensors al- though, among the di erently sized biosensors, these micron-scale biosensors possess the smallest surface area (i.e., the product of the length and width of the biosensor, LW) to cover a Salmonella-inoculated leaf surface. In addition, the LODs for the larger biosensors were found to be higher by half an order to one order of magnitude, 78 and there were not large di erences among them. One possible explanation for these results is that there is a trade-o between the surface area and mass sensitivity of a biosensor as can be seen in Eq. 3.6. In other words, while the surface area of a biosensor increases proportionally to LW, the mass sensitivity decreases proportion- ally to L2WT. Hence, the use of a large biosensor, which covers a large area of a leaf surface, may not always lead to sensitive detection of S. Typhimurium. Rather, a properly small biosensor may be able to show a measurable response even though a small number of S. Typhimurium cells may be bound on this biosensor. In addi- tion, from the preceding microscopic observation, the topography of leaf surfaces and distribution of S. Typhimurium cells are both anticipated to a ect the LOD. Table 4.2: LODs of the di erently sized biosensors for the adaxial and abaxial surfaces of spinach leaves. LOD [cells/cm2] Sensor size Adaxial surface Abaxial surface 2 mm 0.4 mm 15 m 1.28 105 4.12 105 0.5 mm 0.1 mm 15 m 1.42 105 1.37 105 150 m 30 m 4 m 4.77 104 5.22 104 4.4 Discussion 4.4.1 Topography of leaf surfaces and its e ects on the LOD In order to characterize the topography of the leaf surfaces, mean roughness (Ra), mean atness (Fa), and associated major periodicities (TR and TF) were quanti ed. Mean roughness, Ra, is a commonly used roughness parameter and can be de ned as the arithmetic average of the absolute values of height deviations measured from the mean plane of a surface [33, 34]. In addition, the same de nition can be given to mean atness, Fa, which was, however, measured with a longer sampling length (i.e., The sampling lengths for Ra and Fa measurements were 318.5 m and 3185 79 Figure 4.11: Typical height maps (a & b) and associated averaged pro les (c & d) of a leaf surface obtained along di erent sampling lengths. A three-dimensional representation of a leaf surface expressed by Eq. 4.2 is shown in (e). m, respectively.). Figures 4.11a to 4.11d show typical height maps and associated averaged pro les of a leaf surface obtained along di erent sampling lengths. From the pro le with a short sampling length (Fig. 4.11c), it can be seen that the leaf surface possesses a certain degree of roughness, Ra. By contrast, the pro le with a long sampling length (Fig. 4.11d) provides insight into the atness of the surface, Fa. Furthermore, for both surface pro les, the major periodicities, TR and TF, can be determined by performing a Fourier transform of the pro le ordinates and taking the reciprocal of the main frequency mode. The experimentally determined values for these surface geometric parameters are summarized in Table 4.3. The values are all non-zeros, indicating that leaf surfaces are rough, non- at surfaces with topographic periodicities. Hence, with all the geometric parameters, the height at an arbitrary location of a leaf surface, H (x, y), can be expressed by the superposition of sine 80 Table 4.3: Surface geometric parameters for the adaxial and abaxial surfaces of spinach leaves. The values are the averages of 15 samples. Parameter Adaxial surface Abaxial surface Ra ( m) 1.2 0.2 1.6 0.3 Fa ( m) 8.7 3.7 7.0 1.9 TR ( m) 56.3 18.4 67.4 27.0 TF ( m) 1376.7 309.0 874.7 196.1 waves (i.e., when Fa << TF) as H(x;y) = 12[Ra sin( 2 T R x) +Fa sin( 2 T F x) +Ra sin( 2 T R y) +Fa sin( 2 T F y)]; (4.2) where x and y are the lateral coordinates of the leaf surface, respectively. The rst and third terms on the right-hand side of Eq. 4.2 describe roughness pro les, whereas the second and forth terms represent atness pro les along the lateral dimensions of the leaf surface, respectively. A three-dimensional representation of this mathematically expressed surface is shown in Fig. 4.11e, where a periodic arrangement of peaks and valleys can be seen. The periodicities, TR and TF, represent the lateral distances between two adjacent peaks of di erent scale, respectively (i.e., roughness-related small peaks and atness-related large peaks). In order to investigate e ects of a leaf?s surface topography on the physical con- tact between a biosensor and S. Typhimurium cells, the surface area of the biosensor, LW, was compared with two characteristic areas of the leaf surface, AR and AF, which are de ned by AR = TR TR; (4.3) AF = TF TF; (4.4) AR <>AF, the biosensor is likely to stay on large peaks, which greatly reduces the degree of physical contact with the leaf surface (and thus, with S. Typhimurium cells). By contrast, when AF >> LW >> AR, the biosen- sor can t among large peaks (but stays on small peaks), which may improve the physical contact. Furthermore, when LW << AR, the e ects of the leaf?s surface topography can be neglected, and thus, the degree of physical contact can be maxi- mized. In this case, all that matters is the distribution of S. Typhimurium cells on the leaf surface. Hence, the characteristic area, AR, and its associated length, TR, are important parameters that allow one to choose a properly small biosensor free from the surface topographic e ects. As shown in Table 4.3, the values of TR for the adaxial and abaxial surfaces of spinach leaves were 56.3 18.4 m and 67.4 27.0 m, respectively. Among the di erently sized biosensors used in this investigation, the 150- m long sensors possess the closest lateral dimensions to these characteristic lengths, indicating that the topographic e ects for these micron-scale biosensors were the smallest. By contrast, the millimeter-scale biosensors (i.e., 2-mm and 0.5-mm long sensors) possess lateral dimensions much larger than TR and, rather, close to TF. Hence, it is understandable that much larger topographic e ects were posed to these larger biosensors. Furthermore, for real leaf surfaces, there exists microscopic irregularity in topography as can be seen in Figs. 4.11a to 4.11d, which is likely to cause reduced degrees of physical contact particularly for large biosensors. Hence, with the aforementioned merit of high mass sensitivity, the use of su ciently small biosensors (i.e., LW <>> >< >>> >: 1 (N Nmin) 0 (N