MULTICOLOR FLUORESCENCE OPTICAL TWEEZERS METHODS AND APPLICATIONS TO NUCLEIC ACID FOLDING By Cho-Ying Chuang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Physics - Doctor of Philosophy Quantitative Biology - Dual Major 2019 ABSTRACT MULTICOLOR FLUORESCENCE OPTICAL TWEEZERS METHODS AND APPLICATIONS TO NUCLEIC ACID FOLDING By Cho-Ying Chuang Nucleic acids and proteins are fundamental molecular units of life. To understand their properties, we need powerful tools that allow investigation at the single-molecule level. Over the past three decades, the development of single-molecule force and fluorescence tech- niques has provided us new knowledge that was previously unattainable through ensemble measurements. However, we lack methods that allow us to precisely measure the mechanical properties of these molecules while visually detecting multiple molecules at the same time. In this dissertation, we maximize the information obtained in single-molecule measurements by pushing the techniques to be more precise and more complex. We then utilize our instrument to directly observe the folding and unfolding of the nucleic acid G-quadruplex structure. Among the single-molecule force techniques, angstrom-resolution has been achieved by optical tweezers using the dual-trap instrument design. Dual-traps can be produced by acousto-optic (AO) devices, which have many advantages, but trap positioning inaccuracies have limited their usage at high-resolution. We have designed a method to remove the inaccuracies by randomizing the phase of the radio frequency that drives the AO device. We demonstrated that the trap inaccuracies are completely eliminated and high-resolution trapping quality is achieved. This advance allows us to perform long duration measurements with reduced drift in trap measurement over time. Next, we present instrumentation advances that combine high-resolution optical tweez- ers and multicolor confocal fluorescence spectroscopy along with automated single molecule assembly. Multicolor not only allows the detection of multiple observables but also increases the flexibility in the choice of fluorophores. We demonstrated the ability to simultaneously measure angstrom-scale changes in tether extension and single fluorophore signals. The biggest challenge in integrating optical tweezers and fluorescence is the potential for greatly enhanced photobleaching which can make experiments impossible. We showed that the mean number of photons emitted before bleaching is unaffected by the trap laser when interlacing the fluorescence and optical trap lasers. We investigated the photostability of quantum dots and fluorophores. Finally, we devised computer-controlled automation to conserve the fluo- rophore lifetime. This advance enables us to observe multiple molecules or multiple degrees of freedom within a molecular complex while mechanically manipulate and detect them. Taking advantage of these method and instrumentation advances, we investigate the folding and unfolding of a DNA secondary structure: thrombin-binding aptamer G- quadruplex (TBA-GQ). Studying the kinetics of G-quadruplex formation is essential for understanding telomere regulation (the ends of chromosomes) and therapeutic approaches for disease. TBA-GQ is the smallest G-quadruplex. Although many experiments and simula- tions have been done on G-quadruplex, the small size and low stability make it very difficult to observe folding and unfolding of TBA-GQ directly. Our high-resolution optical tweezers have the sensitivity and stability to directly observe TBA-GQ at very low forces. We found that with increasing force, the folding rate decreased and the unfolding rate increased. Our work demonstrates that at a given force, the TBA-GQ formation is facilitated by metal ions and is stabilized by thrombin. It also indicates that the equilibrium force increased as KCl concentration increased. From a detailed analysis of the folding and unfolding rate constants vs applied force, we were able to detect a single transition state conserved across all conditions and identify the structure of the transition state as the G-triplex structure. To my parents Fu-Mei and Ching-Shiang iv ACKNOWLEDGMENTS This dissertation would not have been possible without the guidance and support of many people. I would like to start by thanking my advisor, Professor Matt Comstock, for being supportive and challenging. He gave me the opportunity to enter the field of single-molecule biophysics. I gained invaluable experience on setting up a new lab from scratch. He pro- vided training and resources that guided me through the process of being an independent researcher. I would also like to express my deepest gratitude to Professor Lisa Lapidus for providing insightful suggestions. As a faculty in biophysics, she has been one of my mentors in these years. I’m extremely grateful to my guidance committee members: Professor Matt Comstock, Professor Lisa Lapidus, Professor Heedeok Hong, Professor Carlo Piermarocchi, and Professor Stuart Tessmer, for offering helpful suggestions. I would like to extend my thanks to our collaborator, Professor Hamza Balci (Kent State University), for his inputs in the project. Many thanks to all the past and present members of Comstock Lab and Lapidus Lab, especially to Dena Izadi, Jess West, Miles Whitmore, Matt Zammit, and Drew Baker. The time we spent together was the greatest moment in my graduate school life. I also wish to thank the members in our dissertation writing group, Aimee Shore and Matt Mondragon, for their emotional support and pushing this dissertation over the finish line. I would like to thank the secretaries in the department: Kim Crosslan, Cathy Cords, and Debbie Barratt, for their assistance in both work and life. I very much appreciate for getting hug from Kim whenever I needed it. I would also like to extend my sincere thanks to graduate directors, Professor Scott Pratt and Professor Kirsten Tollefson, and the department chair (now the v dean of College of Natural Science) Professor Phil Duxbury. This work would never have been possible without their advice and support. Thanks also to National Science Foundation, Department of Physics and Astronomy at MSU, and College of Natural Science at MSU, for the financial support. Finally, I’m deeply indebted to my family and friends. Their love helped me build resilience against the stressful life events through these years. My mom Fu-Mei Chiu, my dad Ching-Shiang Chuang, and my sister Po-Han Chuang, have always been there for me through all the ups and downs. My in-laws, Hui-Jen Chien and Ching-Tsai Pan, provided unparalleled support and encouragement. My child, Chensi Pan, brought me endless joy. Last but not least, I would like to express my appreciation to my husband, Kuo-Chuan Pan, for his continued support and the precious memories we share together through the journey. vi TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x Chapter 1 1.1 Nucleic acid molecules Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1.1 Nucleic acid structures . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1.2 Polymer model of unfolded DNA . . . . . . . . . . . . . . . . . . . . 1.1.3 Folding and unfolding: Free energy landscape model . . . . . . . . . . 1.1.4 Beyond Watson-Crick pairs: G-quadruplex structure . . . . . . . . . 1.2 Single-molecule measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.1 Optical tweezers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.2 Fluorescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.3 Optical tweezers + fluorescence microscopy . . . . . . . . . . . . . . . 1.3 Outline of this dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 2 2 4 7 12 18 21 29 33 36 Chapter 2 Randomizing phase to remove acousto-optic device wiggle errors 37 38 40 43 43 45 48 51 53 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.1 AO device positioning inaccuracies . . . . . . . . . . . . . . . . . . . 2.3.2 Untethered beads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.3 Tethered beads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 3 Combined high-resolution optical tweezers and multicolor single- 3.1 3.2 Material and methods molecule fluorescence . . . . . . . . . . . . . . . . . . . . . . . . . . Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1 DNA tether construct preparation . . . . . . . . . . . . . . . . . . . . 3.2.2 Experiment conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.3 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Instrument Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1 Optical tweezers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.2 Multicolor single-molecule fluorescence . . . . . . . . . . . . . . . . . 3.3.3 Bright-field imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.1 Cryptic quantum dot binding for robust signal confocal alignment . . 3.4.2 Multicolor DNA probe binding demonstration . . . . . . . . . . . . . 3.3 54 55 57 57 58 60 61 63 65 71 71 71 77 vii 3.4.3 Cy5 outperformed other organic fluorophores . . . . . . . . . . . . . . 3.4.4 Automated single molecule assembly line . . . . . . . . . . . . . . . . 3.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 83 85 4.1 4.2 Material and methods Chapter 4 Mechanical folding pathway of the thrombin-binding aptamer G-quadruplex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 4.2.1 Dual-trap high-resolution optical tweezers . . . . . . . . . . . . . . . 91 4.2.2 Construct preparation . . . . . . . . . . . . . . . . . . . . . . . . . . 91 4.2.3 Experimental procedure . . . . . . . . . . . . . . . . . . . . . . . . . 92 4.2.4 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 4.3.1 Force dependent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 4.3.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 4.3.3 Thrombin dependent . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 4.4.1 Bell model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 4.4.2 Force-dependent structure diameter model . . . . . . . . . . . . . . . 105 4.4.3 Finding transition state structure . . . . . . . . . . . . . . . . . . . . 107 4.4.4 G-triplex as the transition state . . . . . . . . . . . . . . . . . . . . . 112 4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 Salt dependent APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 Appendix A Protocols for performing experiments . . . . . . . . . . . . . . . . . . 116 Appendix B List of essential parts for the construction of the multi-color fleezers . 128 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 viii LIST OF TABLES Table 4.1: Bell model fit parameters of rate constants and distances to transition state. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 Table 4.2: Fit parameters of rate constants and transition state diameters with a segment of ssDNA (0, 4, and 8 free nucleotides) in 800 mM KCl buffer. 109 Table 4.3: Fit parameters of rate constants and transition state position with a range of KCl concentration using the G-triplex plus 4 free nucleotides model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 Table A.1: Complementary hairpin sequence and probe strand for “cryptic binding” experiments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 Table A.2: ssDNA Insert for synthesizing the “hybridization” construct and the com- plementary fluorophore-labeled probes. . . . . . . . . . . . . . . . . . . 124 Table A.3: ssDNA Insert for synthesizing the “fixed probe” construct and the com- plementary fluorophore-labeled probes. . . . . . . . . . . . . . . . . . . 125 Table A.4: ssDNA Insert for synthesizing the “TBA-GQ” construct. . . . . . . . . 125 ix LIST OF FIGURES Figure 1.1: DNA and RNA are polymer chains. . . . . . . . . . . . . . . . . . . . Figure 1.2: Freely-jointed chain and worm-like chain polymer models. . . . . . . . Figure 1.3: Free energy landscape for a two-state reaction. . . . . . . . . . . . . . . Figure 1.4: Schematic view of how optical tweezers apply forces to the ends of the . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . molecule. Figure 1.5: Effect of the applied force on the free energy landscape for a two-state system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 6 8 10 10 Figure 1.6: Nucleobases are components of nucleotides that form DNA and RNA. . 12 Figure 1.7: Alternative base pairing explains how nucleic acids form a variety of . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . structures. 13 Figure 1.8: G-quadruplex structure. . . . . . . . . . . . . . . . . . . . . . . . . . . 15 Figure 1.9: Schematic structure of thrombin binding aptamer G-quadruplex (TBA- GQ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 Figure 1.10: Biomolecules in action. . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 Figure 1.11: High-resolution optical tweezers combined with confocal microscopy. . 21 Figure 1.12: Ray optics diagrams of optical trapping. . . . . . . . . . . . . . . . . . 23 Figure 1.13: Schematic of the dual-trap optical tweezers setup. . . . . . . . . . . . . 25 Figure 1.14: Principles for acousto-optic devices and electro-optic devices. . . . . . . 26 Figure 1.15: Position detection by back-focal-plane interferometry. . . . . . . . . . . 27 Figure 1.16: Power spectrum of a trapped bead. . . . . . . . . . . . . . . . . . . . . 29 Figure 1.17: Jablonski energy diagram of the fluorescence process. . . . . . . . . . . 30 Figure 1.18: Cy3-Cy5 FRET pair. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 x Figure 1.19: Optics layout of confocal microscope. . . . . . . . . . . . . . . . . . . . 33 Figure 1.20: Interlacing and time-sharing of the optical traps and fluorescence exci- tation lasers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 Figure 2.1: Interference patterns in acousto-optic device. . . . . . . . . . . . . . . . 39 Figure 2.2: Optical tweezers layout. . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 Figure 2.3: RF phase control. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Figure 2.4: Force versus extension measurements pulling on DNA. . . . . . . . . . . 44 Figure 2.5: Wiggles shift over time with acousto-optic device temperature. . . . . . 44 Figure 2.6: Demonstration of positioning and measurement inaccuracies and the cor- . . . responding spatial power spectra for dual-optical tweezers setups. 47 Figure 2.7: Bead measurement noise for coherent versus random phase methods. . . 48 Figure 2.8: Trap laser profile after AOD and the corresponding Fourier power spectra. 50 Figure 3.1: Schematic of the instrument layout. . . . . . . . . . . . . . . . . . . . . 62 Figure 3.2: Interlacing and time-sharing of optical traps and fluorescence excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . lasers. 64 Figure 3.3: Excitation and emission spectra of the fluorescent molecules. . . . . . . 66 Figure 3.4: PSD scan images from all three fluorescence excitation lasers using the piezo mirror stage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 Figure 3.5: Axial alignment of confocal spot. . . . . . . . . . . . . . . . . . . . . . . 70 Figure 3.6: “Cryptic binding” experiments with quantum dots. . . . . . . . . . . . 73 Figure 3.7: Representative time trace of hairpin cryptic binding experiments. . . . . 75 Figure 3.8: Confocal alignment scan images using quantum dots. . . . . . . . . . . 76 Figure 3.9: Multi-color single fluorophore-labeled oligonucleotide hybridization ex- periment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 Figure 3.10: Force dependent DNA extension change upon hybridization. . . . . . . 78 xi Figure 3.11: Simultaneous measurements of fluorescence and tether extension. . . . 79 Figure 3.12: Multi-color single fluorophore-labeled oligonucleotide hybridization ex- . . . . . . . . . . . . . . . . . . . . . . . . periment under 10 pN force. 80 Figure 3.13: Experimental setup for fluorophore lifetime experiments. . . . . . . . . 81 Figure 3.14: Fluorophore lifetime experiments with fixed probes. . . . . . . . . . . . 83 Figure 3.15: Multi-channel flow chamber. . . . . . . . . . . . . . . . . . . . . . . . . 84 Figure 3.16: Molecular assembly line with the multi-channel flow chamber. . . . . . . 85 Figure 4.1: Structure of thrombin-binding aptamer G-quadruplex (TBA-GQ). . . . 90 Figure 4.2: Schematic of the TBA-GQ experimental setup using optical tweezers. . 92 Figure 4.3: Single-molecule optical tweezers study of thrombin-binding aptamer G- quadruplex (TBA-GQ). . . . . . . . . . . . . . . . . . . . . . . . . . . . Figure 4.4: Folding and unfolding dwell times show single exponential distribution. 93 95 Figure 4.5: Equilibrium measurements of TBA-GQ under force. . . . . . . . . . . . 99 Figure 4.6: Force-dependent rate constants of TBA-GQ over a range of KCl concen- tration with Bell model fitting. . . . . . . . . . . . . . . . . . . . . . . . 100 Figure 4.7: Salt effects on TBA-GQ formation. . . . . . . . . . . . . . . . . . . . . 102 Figure 4.8: Thrombin effects on TBA-GQ resolution. . . . . . . . . . . . . . . . . . 104 Figure 4.9: Schematic possible transition state structures during the folding and un- folding pathway of TBA-GQ. . . . . . . . . . . . . . . . . . . . . . . . . 108 Figure 4.10: Force-dependent rate constants of TBA-GQ over a range of KCl concen- tration with force-dependent structure diameter model. . . . . . . . . . 111 Figure 4.11: Proposed reaction diagram explains the folding and unfolding pathway of TBA-GQ. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 Figure A.1: Gel purification. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 xii Chapter 1 Introduction Life comes with various forms. Despite the different appearances on the large scale, all life consists of similar components at the molecular level: nucleic acids (DNA and RNA) and proteins. Nucleic acids and proteins are fundamental biomolecules of life. Therefore, studying the properties of nucleic acids and proteins is essential to unveil the secret of life. All living cells contain nucleic acid molecules that encode the genetic information and utilize the information to synthesize proteins. On the other hand, proteins are the molecular machines that cells use to read the genetic information in nucleic acid molecules. The Central Dogma of Molecular Biology [1] describes how genetic information flows in cells from DNA sequences to RNA transcripts to proteins. The genetic information in DNA is transcribed to messenger RNA (mRNA) molecules, and the mRNA molecules are translated to proteins. All cells in the human body contain the same DNA. Yet the cells are trans- formed into all kinds of different types (e.g. muscle, nerve, skin, etc.). How do the cells know what genes to turn on and when to do which reaction? It is regulated by the nu- cleic acid-protein interactions and the physical conformations of these biomolecules. Beyond protein-encoding genes, nucleic acids play another essential role: regulation. In fact, only a small percentage of the genome is for coding proteins. The vast majority of functional DNA in human genome is for regulation purpose [2]. Some non-coding DNA sequences form unique structures, e.g. G-quadruplex, under certain conditions (see discussion below). Some non-coding DNA is transcribed into non-coding RNA molecules including ribosomal RNA 1 (rRNA), transfer RNA (tRNA), micro RNA (miRNA) etc. They involve in many cellular processes and regulate the gene expression. For example, ribozymes are structures formed by RNA molecules with catalytic properties. Thus, the structure and stability of the nucleic acid and proteins are very important to their function. The more we understand how they work, the more we understand the life. The knowledge is also applicable to curing diseases and creating new technology. 1.1 Nucleic acid molecules 1.1.1 Nucleic acid structures DNA, RNA, and proteins are three fundamental biopolymers. Both DNA and RNA consist of nucleotide chains while proteins consist of amino acid chains. Each nucleotide is composed of three subunits: a nitrogenous base, a 5-carbon sugar (a deoxyribose sugar molecule for DNA and a ribose sugar molecule for RNA), and a phosphate group. The sugar and phosphate group link together by phosphodiester bonds to form DNA or RNA polymer (Fig. 1.1). In the following discussion, we focus on DNA polymers. DNA is built by four different nucleotides with same sugar and phosphate group. The phosphate groups carry negative charges, so DNA is a negatively charged polymer. In addition, DNA strand has polarity. The sugar molecule has carbon that labeled from 1’ to 5’, and the phosphate is linked to the sugar through the 3’ carbon and 5’ carbon. The polarity of the sugar-phosphate backbone defines the 3’ and 5’ directionality in DNA. There are two types of bases for nucleotides: purines (adenine and guanine) and pyrimidines (cytosine, DNA only thymine, and RNA only uracil). The polymers can be held 2 together by hydrogen bonds between bases to form secondary structures. In the primary canonical form, two single-stranded DNA (ssDNA) form DNA double helix. It follows the base paring rules that proposed by Watson and Crick in 1953 [3]: adenine (A) binds to thymine (T), and cytosine (C) binds to guanine (G). In RNA, thymine (T) is replaced by uracil (U) [4]. Figure 1.1: DNA and RNA are polymer chains. (a) The sugar molecule has 5 carbons that labeled from 1’ to 5’. The difference between the deoxyribose sugar in DNA and the ribose sugar is RNA is at the 2’ carbon. Deoxyribose has a hydrogen atom while ribose has a hydroxyl group at the 2’ carbon. (b) Each nucleotide is composed of a base, a sugar, and a phosphate group. The figure here shows a ribose sugar with a base Uracil and a phosphate group. (c) ssDNA polymer chain. The sugar and phosphate group of the nucleotides link together to form DNA polymers. New nucleotides are added on the 3’-OH of the strand. Hydrogen atoms that are bonded to carbon atoms are not drawn. 3 1.1.2 Polymer model of unfolded DNA Nucleic acids and proteins are biological polymers that consist of a sequence of nucleotides or amino acids. They are stiff at short length scales and flexible at long length scales. The behavior of the biological polymers such as nucleic acids and proteins can be described using polymer chain models. Freely-jointed chain model and worm-like chain model are two models we used here. In the freely-jointed chain model (Fig. 1.2a), each segment is a fixed length rigid rod (vector (cid:126)r) and its orientation and position are independent of the other segments. A freely- jointed chain polymer is only flexible between the freely hinged segments. It also assumes the polymer is inextensible and neglect the interactions between segments. Worm-like chain (WLC) model (Fig. 1.2b) describes the behavior of the polymer as a continuously semi-flexible rod. It assumes the polymer is inextensible which means the bonds connecting each segment do not stretch. The stiffness of a worm-like chain polymer corresponds to its persistence length Lp. It is the length over which correlation of the orientation of the tangent vector (cid:126)t(s) is lost. The longer the persistence length, the stiffer the polymer. When the maximum length of the polymer (contour length Lc) is much short than persistence length, Lc << Lp, the polymer behaves like a rigid rod. It has been shown in single-molecule force spectroscopy studies that the force-extension behavior can be described by WLC model [5, 6]. The persistence length is about 50 nm (150 base pairs) for dsDNA, about 1 nm for ssDNA, and about 0.6-0.7 for polypeptides [7–9]. Both WLC and FJC models consider the polymer is inextensible which means the extension of the polymer cannot be stretched beyond its contour length. Since real polymer is extensible, the elastic response from the extension needs to be taken into account [10, 11]. 4 We use the modified version: extensible WLC (XWLC) and extensible FJC (XFJC), to predict and fit our experimental data. The polymer model for entire tether is the sum of each component (dsDNA handles and ssDNA inserts). Figure 1.2c shows an example force-extension curve of a single hairpin construct measured by our high-resolution optical tweezers. The DNA hairpin construct is well- described by two dsDNA XWLC polymer handles plus one ssDNA hairpin XFJC polymer insert. The length of the entire tether changes when the hairpin unfolds (increases in exten- sion) and folds (decreases in extension) with the corresponding force changes. The size of the transition corresponds to the length change of the insert. We can calculate force-extension changes based on the number of nucleotides involved in the process. 5 Figure 1.2: Freely-jointed chain and worm-like chain polymer models. (a) Freely- jointed chain polymer model considers each DNA segment as a rigid rod. (cid:126)r is the vector of each segment, and (cid:126)R is the total end-to-end vector. (b) Worm-like chain polymer model considers DNA as a continuously flexible rod. (cid:126)t(s) is the tangent vector along the chain at point s. (c) Force-extension curve of a DNA hairpin tether. The folding and unfolding behavior of a single hairpin construct during pulling (blue) and relaxing (green) is well- described by the folded hairpin (red dashed) and unfolded hairpin (black dashed) polymer models. 6 1.1.3 Folding and unfolding: Free energy landscape model Thermodynamics Nucleic acids participate in almost every function in the living cells. Under certain conditions, the nucleic acids transformed from the unfolded polymer chain of nucleotides into a folded structure. It is fundamental to understand how nucleic acids fold because the mechanical properties are critical to their biological function. The folding and unfolding dynamics of nucleic acids as well as proteins can be de- scribed as the Gibbs free energy landscape of the molecule (Fig. 1.3). Gibbs free energy landscape theory provides a way to understand the structure formation of biopolymers. For a particular reaction, each structure of the molecule in the progress, which is the macrostates of the molecule, are represented as sets of microstates sharing the same property (i.e. the end-to-end extension in this dissertation). The landscape consists of a surface representing the free energy of the macrostates (e.g., folded state, unfolded state and transition state) along the reaction coordinate. The molecules diffuse on the landscape and tend to stay at a state with the minimum free energy, which is typically the folded state. The changes of free energy from macrostate A to macrostate B at constant temperature is, ∆G = ∆H − T ∆S. (1.1) ∆H is the change in enthalpy of the molecule, T is the absolute temperature, and ∆S is the change in entropy of the molecule. The enthalpy of the molecule is from the electrostatic bonding energy between different parts of the molecule, e.g., the hydrogen bonds. The en- thalpy includes the intrachain interaction in nucleic acids or proteins and the intermolecular interaction with surrounding molecules such as water and ions. The entropy is derived from 7 the quantity Ω which represents the number of possible conformational states (microstates) of the molecule as where kB is the Boltzmann constant. S = kB · ln Ω (1.2) Figure 1.3: Free energy landscape for a two-state reaction. ‡ indicates the transition state. Here, we describe the free energy surface in a two-state system where the molecule is in either the folded state or the unfolded state with an energy barrier, transition state (indi- cated by ‡), in between. In general, the folded state is a very specific structural conformation of the molecule with a maximum number of intramolecular noncovalent bonds. As such, for the folded state, there are very few conformation states. Thus the folded state has a minimal entropy. However, for the unfolded state, the unfolded state polymer has the largest number of possible conformational states because the molecule is a polymer chain with no bonds between subunits there are many random microstates with the same end-to-end extension. Therefore, the unfolded state has the highest conformational entropy. If we add one bond, 8 the molecule moves to the left on the diagram because the end-to-end extension is shortened. The enthalpy is lower by the addition of the bond while the entropy is also lower due to the reduced number of microstates. The enthalpy term and the entropy term of the free energy of the molecule compete with each other resulting a free energy landscape along the reaction coordinate. Generally, when we start adding bonds, the entropy goes down more than the enthalpy does, so the free energy is a net increase. At some point, the molecule reaches the top of the barrier (transition state). Adding bonds start to outweigh the loss of entropy, and the free energy starts goes down. Eventually, the molecule arrives at the folded state. Force-dependent free energy landscape In the absence of force, the minimum free energy for biological molecules is generally the folded state and thus the molecules are most likely found folded. When applying a force to the ends of a molecule (Fig. 1.4), work is done on the molecule. It changes the enthalpy of the molecule, thus the free energy changes. The force tilts the energy landscape along the mechanical reaction coordinate by a factor F∆x where ∆x is the distance between the point of interest and the reference point. As the force increases, the unfolded state becomes more and more energetically favorable. At equivalent force F1/2, the change in free energy is zero, the molecule has equal probability to stay in either the folded state or unfolded state. When the applied force is larger than equivalent force, the reaction favors unfolded state (Fig. 1.5). 9 Figure 1.4: Schematic view of how optical tweezers apply forces to the ends of the molecule. The introduction of optical tweezers will be given in Section 1.2.1. Figure 1.5: Effect of the applied force on the free energy landscape for a two- state system. The red curve is the free energy surface at zero force, the green curve is at equivalent force F1/2, and the blue curve is at the force higher than F1/2. With the applied force, the relative free energy differences between the transition state and folded (unfolded) state are changed. Curves are vertically shifted for clarity. 10 Transition state theory provides an approach to describe the rate constants of the reaction. The rate constant of reaction depends on the free energy and the temperature. k0 = A exp(−∆G‡ kBT ) (1.3) A is the intrinsic rate constant for the system (like an attempt frequency), and ∆G‡ is the activation energy (barrier height) for unfolding or folding. The simplest Bell model describes how the applied force affect the barrier height. In the presence of force F, the Gibbs free energy will be lower by F∆x. The activation energy becomes ∆G‡ - F∆x. Because what is limiting is overcoming the energy barrier, all that matters for the effect of force is the distance to the transition state. Therefore, the force dependent rate constant become k = A exp(−∆G‡ kBT ) exp( F ∆x kBT ) = k0 exp( F ∆x kBT ) (1.4) where k0 is the rate constant at zero force. The force affects the energy landscape by changing the relative barrier height to the transition state, and thus the rate constants of reaction are affected. By measuring the force dependent rate constants, we can obtain the distance to transition state from folded and unfolded states and furthermore reconstruct energy landscape along the mechanical reaction coordinate. The energy landscape allows us to characterize the mechanical properties of biopolymers. 11 1.1.4 Beyond Watson-Crick pairs: G-quadruplex structure The mechanical properties of the biopolymers depend on the binding between subunits. Although Watson-Crick base pairs are the most abundant base pairs in nucleic acids, a nucleotide has more than one edge available for hydrogen bonding [12]. Figure 1.6 shows the Watson-Crick edge and Hoogsteen edge of the nucleobases. Through the Watson-Crick base pairs and Hoogsteen base pairs (observed in 1963) [13], nucleic acids can form alternative nucleic acid structures (Fig. 1.7) like stem loop, pseudoknot, triplex and G-quadruplex. Figure 1.6: Nucleobases are components of nucleotides that form DNA and RNA. Depending on the edges of the bases involved in the hydrogen-bonded base pairing, DNA and RNA form canonical base pairs with Watson-Crick edge (blus solid curves) and non-canonical base pairs with Hoogsteen edge (red dashed curves). R represents the sugar-phosphate backbone. 12 Figure 1.7: Alternative base pairing explains how nucleic acids form a variety of structures. (a) DNA duplex. Watson-Crick base pairs (blue dashed lines) are the building blocks of the most common double-stranded DNA (dsDNA) helix. T is paired with A via two hydrogen bonds. C is paired with G via three hydrogen bonds. (b) DNA triplex. One DNA strand binds to the dsDNA helix via Hoogsteen base pairs (red dashed lines) form triple- stranded DNA. In triple-helix base pairing, Watson-Crick base pair is indicated by -, and Hoogsteen base pair is indicated by :. (c) G-quadruplex. Four G bases form G-tetrad via 16 Hoogsteen base pairs. Two or more G-tetrads stack on top of each other form G-quadruplex. 13 G-quadruplex G-quadruplex is formed by guanine-rich single-stranded nucleic acids via Hoogsteen base- pairing [14]. Four coplanar guanines that held together by 8 Hoogsteen hydrogen bonds and associate with a monovalent metal ion such as Na+ and K+ to form G-tetrad (Fig. 1.8a), the building block of G-quadruplex. Two or more G-tetrads stack on top of one another form G-quadruplex. Recent studies have indicated that G-quadruplex structures play cru- cial regulatory roles in replication, transcription, and translation [15]. During these bio- logical processes, the double-stranded DNA (dsDNA) becomes single-stranded that permits the guanine-rich sequence to form G-quadruplex. The highest abundance of guanine-rich se- quences is at telomeres, the 3’ single-stranded DNA tail at end of the eukaryote chromosomes. Telomeres protect the linear ends of eukaryote chromosomes. The telomeric G-quadruplex inhibits the activity of telomerase, a reverse transcriptase which is active in cancer cells. Since G-quadruplex is under mechanical force before interaction with proteins, investigation of the folding mechanism of G-quadruplex under force is important for molecular biology understanding and structure-based drug designs. The types of G-quadruplex structure can be categorized based on the number of the nucleic acid strands involved (intramolecular or intermolecular) (Fig. 1.8b) and the orientation of strands (parallel, hybrid, or antiparallel) (Fig. 1.8c). The intramolecular G- quadruplex is formed by a single-strand guanine-rich nucleic acids, while the inter molecular G-quadruplex consists of two or four separate strands. In parallel type of G-quadruplex structure, the direction of all strands is the same. For example, all strands are aligned from 5’ to 3’ direction. In hybrid type, one of the runs of guanine bases has the opposite orientation to the other three runs. In antiparallel type, two runs have opposite orienta- tion to the other two runs. The conformation of the G-quadruplex is influenced by the 14 nucleic acid sequence, length, and the environmental conditions such as cation species and molecular crowding [16–21]. Even very similar sequences can produce different conformation. Figure 1.8: G-quadruplex structure. (b) Examples of intermolecu- lar G-quadruplex (tetramolecular and bimolecular) and intramolecular G-quadruplex (uni- molecular). Strand polarities are from 5 to 3. (c) Possible geometries of intramolecular G-quadruplex. (a) G-tetrad. Thrombin binding aptamer G-quadruplex The G-quadruplex structures has been adopted by several aptamers. Aptamers are short DNA or RNA strands that bind to target molecule with high affinity and specificity via their unique folded structures. They are selected from a large pool of sequences through poly- merase chain reaction (PCR) process, known as systematic evolution of ligands by exponen- 15 tial enrichment (SELEX) [22–24]. G-quadruplex forming aptamer is one of the most popular categories among the nucleic acid aptamers. Even with similar sequences, G-quadruplex forming aptamers are able to produce a large variety of conformations, which make them recognize a wide range of protein targets. The identification of the target is critical in drug discovery. Compared to the protein-based antibodies, nucleic acid-based aptamers have the advantages because of their smaller size, lower immunogenicity, low cost, and easiness to synthesize [25]. It can be widely used for therapeutics, diagnostics, biosensor, and molecular biology studies. Thrombin binding aptamer G-quadruplex (TBA-GQ) is a minimal model system of G- quadruplex [26]. Its a 15-nucleotide long single-stranded DNA that folds into a chair-like two- layer G-quadruplex with one TGT loop and two TT connecting loops (Fig. 1.9). Because it is the simplest G-quadruplex, understanding the mechanical properties of TBA-GQ is desired to further investigate the mechanism of other important G-quadruplex structures such as telomeric G-quadruplex. Figure 1.9: Schematic structure of thrombin binding aptamer G-quadruplex (TBA-GQ). TBA-GQ is an intramolecular, antiparallel G-quadruplex with a chair-like conformation. TBA-GQ is also an important DNA aptamer on its own. The protein thrombin is 16 the central enzyme in the process of blood clotting. The regulation of thrombin activity is important for anticoagulant and cardiovascular diseases therapy. TBA-GQ specifically interacts with thrombin to inhibit the blood coagulation. As a result, TBA-GQ has been tested in phase I clinical trials that known as ARC183 or HD1 (Archemix Corporation) [27]. G-quadruplex is functionally important in genome. To understand how G-quadruplex works in cells, we need to understand the structure stability (i.e. the folding and unfold- ing mechanism). TBA-GQ is a small DNA structure. Compared to other G-quadruplex structures, it only has two layers that means it contains fewer bonds thus less stable. The small size and low stability make it very technically challenging to directly observe the fold- ing/unfolding of TBA-GQ. As a consequence, a high-resolution measurement is needed to study TBA-GQ formation and thrombin-aptamer interaction. 17 1.2 Single-molecule measurements To comprehensively study the structure and function of the biomolecules as well as the inter- actions between them, we need powerful tools that allow investigation at the single-molecule level. It is technically challenging to study these small but fundamental biomolecules indi- vidually. Over the last three decades, single-molecule force and fluorescence techniques have been developed. These techniques have been used to characterize the mechanical properties and the basic mechanisms of the DNA, RNA, and proteins. Single-molecule methods characterize the properties of individual molecules. In the past and to this day, the majority of molecular biology experiments are ensemble-based methods. These experiments measure the average properties of molecules. The information we can learn from the average properties is limited because the details in the biological sys- tems are dynamics. For example, the ensemble measurements can give us the ratio of folded to unfolded state in a group of molecules, but we are not able to observe how a molecule changes its configuration between unfolded state to folded state. In contrast, single-molecule methods [28] provide information about the biological systems that was unattainable through the ensemble experiments. Without the averaging of time and population, single-molecule methods not only provide the mean values of the system but also the dynamics and nonuni- form behavior such as the transient states. Force-based and fluorescence-based are two major branches of the single-molecule methods. Single-molecule force spectroscopy is a powerful tool to study the mechanical properties and the dynamics of biomolecules such as protein folding, nucleic acid folding, and the nucleic acid processing molecular motors. Atomic force microscopy (AFM) [29], optical tweezers [30], and magnetic tweezers [31] are three commonly-used single-molecule force spectroscopy. Force spectroscopy can manipulate in- 18 dividual molecules as well as quantitatively apply and measure the force on the biological system. Single-molecule fluorescence microscopy uses fluorophores conjugated to individual biomolecules as a reporter in a wide variety of measurements including the detection of the presence, location and/or the conformation of the molecules of interest. The commonly used single-molecule fluorescence microscopy includes confocal microscopy [32], total internal re- flection microscopy (TIRF) [33], light sheet microscopy [34], and super resolution microscopy [35]. Each technique has its own strengths and limitations. It is important to push the instrument capabilities to more complexity and higher precision to maximize the informa- tion and details obtained. Typical single-molecule techniques mentioned above measure a single quantity, e.g. the displacement via optical tweezers or the location via fluorescence. However, biomolecules are complex. A reaction requires many components and involves many degrees of freedom (Fig. 1.10). Typical single-molecule techniques are not sufficient to understand the complexity of biology. The development of single-molecule techniques has been moving towards the combination of multiple techniques to directly probe the complex biological systems. By combining force and fluorescence, we can visualize the single molecule fluorescence while mechanically manipulate and detect the molecules. Here we describe the principles of optical tweezers and fluorescence, and the hybrid technique (Fig. 1.11) that combining both, the main technique we use in this dissertation. 19 Figure 1.10: Biomolecules in action. (a) The telomeric DNA replication fork. (b) The coordination of multiple proteins is required for telomeric DNA replication, and multi-degree of freedom conformational changes in DNA and proteins are essential to function properly. See Reference [36] for details. Figure is reused under permission of a Creative Commons license (CC BY 4.0). 20 Figure 1.11: High-resolution optical tweezers combined with confocal microscopy. Combining both force and fluorescence approaches provide complementary information. 1.2.1 Optical tweezers How do we trap? Optical tweezers are a technique utilizing a laser to trap and manipulate objects [37, 38]. The trapped objects need to be transparent (no absorption) and have the refractive index higher than the medium. A trapped object senses a net force due to the momentum transfer from the scattering and refracting photons. The forces exerted by light on an object can be divided into two categories: the scattering force and the gradient force. Depends on the size of the trapped objects and the wavelength of the trapping beam, there are two models to quantify the forces: the dipole model and the ray optics model. For a trapped object with 21 diameter d much smaller than the wavelength λ (d << λ), the object can be considered as a point dipole in an inhomogeneous electromagnetic field. The restoring force pushes the dipole to the focus of the trapping beam which is the location of the highest field. For a trapped object with diameter d much larger than the wavelength λ (d >> λ), the forces can be explained using ray optics (Fig. 1.12). For the intermediate regime (d ∼ λ)), which is our case (800 nm bead and 1064 nm trap laser), behavior of a trapped object is the combination of these two models. When the incident light hits the object, some photons are absorbed and some pho- tons are reflected on the surface. The net scattering force pushes the object forward in the direction of light propagation. The gradient force results from the refraction of light between the object and the surrounding medium as the light passes through the object. What mat- ters most is the rays at the edge because they have bigger refraction angles. In the lateral direction, the object experiences a light gradient because the trapping beam has a Gaussian beam profile. The resulting gradient force pulls the object toward the center where the intensity of the light is the highest. In the axial direction, the scattering force pushes the object out of focus as previously described. The axial gradient force needs to pull the object back to compensate the scattering force so that the object can be stably trapped. As shown in Fig. 1.12d, a strong axial intensity gradient can be created by tightly focusing the beam using a high numerical aperture microscope objective. The net force pulls the object to the focus of the optical traps. 22 Figure 1.12: Ray optics diagrams of optical trapping. The momentum transfer occurs when a beam of light (orange arrows) hits a dielectric bead (grey spheres). (a) Scattering forces. The incident light is reflected on the bead surface and results in the momentum changes of photons (green arrows). The net force pushes the bead forward in the direction of light propagation. (b) Lateral gradient force. The bead experiences an intensity gradient (orange bars) from the incident light. The light is refracted from the bead and results in the momentum changes. The net force pulls the bead to the center of the beam. (c) Axial gradient force. The net force pulls the bead toward the highest intensity. (d) Axial gradient force with a high numerical aperture lens. The incident light comes in at large angles. The net force pulls the bead to the focal point. 23 How do we manipulate? One of the advantages of optical tweezers in biology is the ability to manipulate the biological objects. The position of the optical traps can be controlled by 1) moving the sample chamber with respect to the traps or 2) steering the trapping beam itself. The first approach provides large scale (tens of millimeters) distance changes. It can be achieved by using a 3-axis motorized stage. The second approach provides sub-nanometer scale distance changes. It can be done by changing the incoming angle of the trapping beam using piezo scanning mirrors, acousto-optic devices (AO devices) or electro-optic devices (EO devices). The biomolecules such as nucleic acids and proteins are too small to be trapped. Instead, we trap and manipulate the micron-sized dielectric particles (polystyrene or silica beads) that are attached to the molecules of interest. To apply the force, we need the other end of the tether to be fixed. For early generation single optical trap instruments, the other end is either attached to the sample chamber or a bead attached by the suction of a micropipette. In this configuration, the measurement spatial resolution is limited by the drift of sample stage and surface. Recently, high-resolution optical tweezers has been achieved by the dual-trap design [39]. Two traps are generated by rapidly switching the laser between two trap positions. Therefore, the drift of sample stage and surface is removed by isolating the tether from the sample chamber. Moreover, because the dual traps are created from the same laser source and share an almost identical optical path, the fluctuations are reduced by differential detection of the two trapped beads. For a typical dual-trap optical tweezers experimental setup (Fig. 1.13), the molecule of interest is tethered between two trapped beads via long DNA handles. By increasing trap separation, the force is applied to the molecule. When the structure of molecule changes (e.g. unfolding DNA or protein), the tether extension changes and thus the positions of the bead move. By measuring the bead 24 position versus time, we are actually recording the biomolecule configuration versus time. The dual trap are generated by steering the light via AO devices [40] or EO devices [41]. The time-sharing switch rate needs to be high enough to reduce the effect on bead positions. Both devices can rapidly switch the beam between trap positions. The schematic operating principles of AO devices and EO devices are shown in Fig. 1.14. AO devices contain an acousto-optic crystal with a piezoelectric element at the end. When applying a RF signal to the piezo, the vibrating piezo generates acoustic waves in the crystal act like the diffraction grating and deflect the trap laser light. EO devices contain an electro-optic crys- tal. When an electric field is applied, the refractive index changes and the light is deflected. The refractive index is linearly proportional to the applied electric field. AO devices have been used extensively in optical tweezers because of their exibility and speed. However, the AO devices have trap positioning inaccuracies, likely caused by the standing acoustic waves, that limit their usefulness especially for high-resolution applications. EO devices have been used to avoid the errors from AO devices [42], but the deflection range of EO devices are 10 times less than AO devices. In Chapter 2, we discuss a method we devised to remove these inaccuracies and allows AO devices to be used in high-precision measurements. Figure 1.13: Schematic of the dual-trap optical tweezers setup. Two beads (grey spheres) are held in optical traps (orange cones). A molecule of interest (cyan) is tethered between two beads via two long DNA handles (red). The figure is not drawn to scale. 25 Figure 1.14: Principles for acousto-optic devices and electro-optic devices. (a) Acousto-optic device. The incoming beam is diffracted by the RF driven acoustic waves. The angle of the diffracted beam is controlled by the RF frequency. The first-order diffracted outgoing beams cause by two different frequencies (f1 and f2) are shown. (b) Electro-optic device. The incoming beam is deflected by different amounts with the control of applied voltage. The refractive index of electro-optic crystal is changed with the applied electric field. The deflected outgoing beams cause by two different refractive indexes (n1 and n2) are shown. How do we detect? Optical tweezers are not only a manipulation tool but also used for quantitative measurement. To quantitatively detect the displacement of a trapped object and the applied force, several methods have been developed such as video-based [43], direct imaging [38], and laser-based detection [44]. One method has become the standard for the position detection is the back- focal-plane interferometry [45]. Back-focal-plane interferometry uses the interference pattern produced by the forward-scattered light from the trapped object and the unscattered light from the trapping beam. Any displacement of the trapped object results in a change of the light pattern. The interference pattern is monitored by a position sensitive photodetector (Fig. 1.15). The pattern can be calibrated to yield the location of the trapped object. 26 Figure 1.15: Position detection by back-focal-plane interferometry. The interference pattern of a trapped bead is detected by a position sensitive photodetector. The bead is (a) at its equilibrium position, (b) displaced by dy, and (c) displaced by dz. Changes in all three dimensions can be detected. The output signals from the detector are voltage readings. To convert the voltage readings to real displacements and forces, we need to know the conversion factor α which converts the photodetector voltage output to the displacement of the trapped object x = αV , and stiffness κ of the trap which determines the force F = −κx. Calibration can be done with a known displacement or a known force. In our lab, we use Brownian motion as a test force for bead displacement calibration. The equation of motion for the bead can be written as γ dx dt + κx = F (t), (1.5) where x is the position of the bead, γ = 3πηd is the hydrodynamic drag coefficient of the object (for Stokes drag on a sphere of diameter d in a medium with viscosity η), and κ is the trap stiffness. The hydrodynamic drag force plus the trap force equals to the fluctuating Brownian force. 27 Because the average value of the thermal force is zero, a Lorentzian power spectrum Sxx(f ) = | ˜F (f )|2 = kBT c + f 2) π2γ(f 2 (1.6) can be obtained by taking the Fourier transform of the force F(t). Sxx(f ) is in unit of nm2/Hz, kB is Boltzmanns constant, T is the absolute temperature, and fc = κ 2πγ is the corner frequency with the trap stiffness κ. The signals from detector are calibrated by measuring the power spectrum of the Brownian motion of a trapped bead [40]. The power spectrum measured by the detector is Svv(f ) in unit of Volt2/Hz. It is related to the true power spectrum by Svv(f ) = Sxx(f ) α2 , (1.7) where α represents the linear conversion between the detector output and bead displacement in unit of nm/Volt. For low frequency f << fc, Svv(f ) approaches the limit 4kBT γ κ2α2 . For high kBT πγf 2α2 . By fitting the power spectrum (Fig. 1.16), we frequency f >> fc, Svv(f ) becomes obtain the conversion factor α and stiffness κ. For a trapped object, the force and displacement follow the Hooke’s law F = −κx. (1.8) Once the trap stiffness is calibrated, the readout of the detector can be converted to the displacement of the object and the force. The typical trap stiffness is on the order of 0.1 pN/nm. With these techniques, the resolution of optical tweezers has been pushed to mil- lisecond time scale, sub-nanometer spatial scale, and sub-piconewton force scale. 28 Figure 1.16: Power spectrum of a trapped bead. The raw power spectrum (blue) of an 800 nm polystyrene bead from our optical tweezers and the fit (black) to a Lorentzian. The spectrum is asymptotic to the red dashed straight lines at high frequencies and low frequencies. 1.2.2 Fluorescence Photobleaching Fluorescence is a process of emitting a photon via electronic state transitions. Fluores- cence occurs when the electron of a fluorescent molecule relaxes to its ground state from an excited state via emitting photon. By specifically labeling the molecule of interests with fluo- rescence molecules (for example, organic fluorophores, inorganic quantum dots or fluorescent proteins), we can localize and track the dynamics of the target molecules. The fluorescence process (Fig. 1.17) starts by exciting the fluorescent molecule through absorbing the light at a resonance wavelength. The fluorophore is unstable at high energy configuration, so it de- 29 cays to the lowest energy excited state via vibrational relaxation with a loss of energy. From the semi-stable excited state, the electron then decays to the ground state by emitting the light with longer wavelength. The wavelength of an emitted light is always longer than the absorbed light due to the energy loss in the vibrational relaxation. Because the wavelength is longer, that emitted light can be separated from the much brighter excitation laser light via filters. The fluorescence process can cycle multiple times, emitting many photons, until the fluorophore permanently enters the dark state. When a fluorophore is no longer be able to fluoresce, it is called photobleaching. By counting the photobleaching events, the number of individual labeled molecules presence in a complex is determined. The applications are discussed in Chapter 3. Figure 1.17: Jablonski energy diagram of the fluorescence process. Step 1) Ab- sorption. An absorbed light (green) excites the electron from ground state to excited state. Step 2) Relaxation. The electron decays to the lowest energy excited state via vibrational relaxation. Step 3) Emission. The electron emits a longer wavelength light (yellow-green) and returns to the ground state. 30 FRET Two fluorophores with different colors not only provide us the ability to observe two kinds of molecules, but also provide distance information. Two-color detection measures the dynam- ics and relative distance between a pair of fluorophores, donor and acceptor, through the F¨orster resonance energy transfer (FRET). The emission spectrum of the donor (the energy- giving molecule) overlaps with the excitation spectrum of the acceptor (the energy-receiving molecule) (Fig. 1.18a). The excitation of the donor results in the emission of the acceptor. The energy is transferred by nonradiative dipole-dipole interaction. The FRET efficiency, E = 1 1 + ( R R0 )6 = IA ID + IA (1.9) is distance dependent. R is the distance between donor and acceptor, R0 is the F¨orster radius at which energy transfer FRET efficiency is 50% (a property of a specific pair of fluorophore molecules), and IA and ID are the fluorescence emission intensity of the acceptor and the donor, respectively. One of the most commonly used fluorophore pair for FRET is the Cy3- Cy5 pair. Figure 1.18b shows the distance dependence of FRET efficiency using Cy3-Cy5 pair as an example. R0 for Cy3-Cy5 is about 6 nm. FRET is sensitive at the distance 1-10 nm, giving distance measurements way beyond the diffraction limit of the optical microscope direct imaging. The range is comparable to the dimensions of biomolecules, so it has been used as a spectroscopic ruler [46] to investigate the structural changes of proteins and the relative motion between two biomolecules. 31 Figure 1.18: Cy3-Cy5 FRET pair. (a) The excitation (dashed lines) and emission (solid lines) spectrum of Cy3 (donor, green) and Cy5 (acceptor, red). The spectral overlap between the emission of Cy3 and the excitation of Cy5 makes Cy3-Cy5 an ideal pair for FRET study. (b) Distance dependence of FRET efficiency. When the distance between donor (green disk) and acceptor (red disk) equal to R0 (about 6 nm for Cy3-Cy5 pair), the donor transfers half of its energy to the acceptor. Single-molecule fluorescence confocal microscopy The single-molecule fluorescence detection can be realized for a variety of microscope de- signs. Individual photons are collected and counted with rates only on the kHz scale. Thus, in all cases, great care must be taken in design to minimize background detection and maxi- mize signal detection. Here we introduce the confocal microscopy. In a confocal microscope (Fig. 1.19), the excitation beam is reflected by a longpass dichroic mirror and focused in the focal plane of the objective to a minimal nearly diffraction limited volume in the sample. The emission signals are collected by the same objective, passed through the longpass dichroic mirror, passed through the pinhole, and recorded by a single photon counting avalanche pho- todiode (APD). To obtain a 2D image, a raster scan is made, where fluorescence is collected in sequence for each point in the image. The pinhole is located in the conjugated focal plane. It acts like a “spatial filter” so that the out-of-focus fluorescence signals are blocked by the 32 pinhole. Due to the elimination of out-of-focus fluorescence background, confocal microscope has been widely used in biological imaging. Figure 1.19: Optics layout of confocal microscope. Single channel excitation (green) and emission (yellow-green) are shown. A dichroic mirror (D) spectrally separate lights by reflecting excitation light and transmitting emission light. The filter (F) further blocks the exceed excitation light. 1.2.3 Optical tweezers + fluorescence microscopy Force-fluorescence approaches Both optical tweezers and single-molecule fluorescence microscopy have been widely used to study biological systems, but each technique has its strengths and drawbacks. Optical tweezers provide direct detection of the molecular motion, but all motions are projected onto one single axis. It lacks the ability to observe the internal changes in the system. In con- trast, fluorescence can probe changes within the complex but no mechanical manipulation. These limitations motivate the development of the hybrid techniques. The biggest challenge for implementing optical tweezers with fluorescence microscopy is the photobleaching. The 33 fluorescence signal from a single fluorophore molecule is already very dim. Optical tweez- ers use a high-intensity trapping laser, > 1 watt, while fluorescent excitation intensity at a single fluorophore is typically ∼1 micro watt. The presence of the optical trapping beam not only makes it harder to detect single molecule fluorescence (contributing background photons) but also greatly speeds up photobleaching [47]. One way to tackle this issue is separating the trapping laser and the fluorescently labeled molecules in space [48] to min- imize the impact from trap beam. This approach requires long spacers so it comes at the cost of the instrument resolution (longer spacers are softer and do not transmit signals as well). Another strategy is to separate the trapping and fluorescence excitation beams in time by interlacing [49]. Using the AO devices or EO devices described above, both light sources can be switched rapidly on and off out of phase from each other. This approach minimizes the photobleaching by preventing the fluorophores expose to both light sources at the same time. Moreover, neither optical tweezers spatial resolution nor fluorescence detec- tion sensitivity is sacrificed. By time-sharing and interlacing between trap and fluorescence beams (Fig. 1.20), a high-resolution optical tweezers with single-color confocal microscope has been realized [50]. The simultaneous measurement of sub-nm mechanical changes and single-molecule fluorescence detection has been demonstrated. Since then, similar instru- ments combining optical tweezers and single-color fluorescence excitation have been utilized to address the questions that is difficult to answer by either technique alone [51–54]. Three colors and more The functions and interactions in the biological systems are often very complex that require coordination of multiple proteins and multi-degree of freedom conformational changes. The single excitation laser with two-color FRET limits its ability to observe single species or 34 Figure 1.20: Interlacing and time-sharing of the optical traps and fluorescence excitation lasers. Two optical traps (orange) are created in sequence. The trap laser is switched off while the fluorescence excitation laser (green) is turned on. one pair of relative motion. It also limits the choices of fluorescence molecules. The ability to observe multiple fluorescently labeled molecules and get stable signals are necessary to resolve the detail mechanism. In Chapter 3, we implement the optical tweezers with mul- ticolor fluorescence. We demonstrate the ability to probe three different colors at the same time while mechanically manipulate the biological systems. The photostability of organic fluorophores and inorganic quantum dots are evaluated in this hybrid instrument. We also devised computer-controlled single-molecule assembly line which enables the precise assem- bly of multi-molecule complexes while preserving fluorophores. 35 1.3 Outline of this dissertation Three projects are presented in this dissertation. The main focus of these projects is to address the fundamental questions in biology, the mechanical properties of DNA, using a newly developed high-resolution hybrid single-molecule technique. The method and tool we developed fulfill the need to study individual complex that contains multiple components and multi-degree of freedom at base-pair resolution. Therefore, it can be broadly applied to investigate a wide range of molecular systems such as nucleic acid folding, protein folding, and the interactions between nucleic acids and protein molecular machines. This dissertation contains two major parts. The first part (Chapter 2 and Chap- ter 3) describes the single-molecule instrumentation advances we developed. We present a hybrid single-molecule technique that combines high-resolution optical tweezers with multi- color confocal fluorescence microscopy. We discuss the advances in optical tweezers and the fluorescence performance that increase the instrument accuracy and capability. In the sec- ond part (Chapter 4), we utilize our technique to investigate the conformational stability of the nucleic acid secondary structure, G-quadruplex. 36 Chapter 2 Randomizing phase to remove acousto-optic device wiggle errors∗ Acousto-optic (AO) devices have been used extensively in optical tweezers because of their flexibility and speed; however, these devices have trap positioning inaccuracies that limit their usefulness, especially for high-resolution applications. We show that these inaccura- cies are due to interference patterns within the AO device sound fields. We have devised a method that removes these inaccuracies by reducing the coherence of the sound fields by directly controlling and randomizing the phase of the radio frequency voltage input signal. We demonstrate that the trapping inaccuracies are eliminated, for both constant trap po- sition and force-ramp measurements, and that no additional noise is added. We show that this random phase method is applicable to both acousto-optic modulator and deflector type devices and can be easily integrated via software upgrade into existing instruments. ∗The work in this chapter has been published as [55]: A. G. Baker, Cho-Ying Chuang, M. L. Whitmore, and M. J. Comstock, Applied Optics, 57, 1752-1756 (2018) 37 2.1 Introduction Optical trapping has been used to study molecular machines, cells, and other physical and biological phenomenon at the single particle level and recently at sub-nm spatial resolution [39,40,56,57]. Utilizing a high numerical aperture objective, a powerful laser can be focused to a spot and used to trap transparent dielectric micrometer-sized objects such as polystyrene beads, bacteria and viruses [40]. In single molecule studies, a trapped bead is commonly tethered to a sample chamber surface or a second trapped bead by custom-designed single strands of DNA. The tether is designed so that changes in the tether length over the course of an experiment reveal, for example, the step-wise motion of a motor protein or unfolding of a protein. The positioning of the optical traps via deflection of the incoming trapping laser applies tension to the tether and is used for common methods such as force ramp measurements and active force feedback. High-resolution optical tweezers require precise and stable trap laser deflection con- trol. High stability piezo mirror stages and AO devices have been used previously for this purpose. AO devices deflect lasers by diffracting the laser beam from a periodic grating, created by sound waves traveling in a crystal (Fig. 2.1). This sound wave is generated by a piezoelectric element glued to the crystal and driven by an radio frequency (RF) volt- age signal. AO trap positioners have many advantages including high speed (> 100 kHz); all-electronic control enabling, e.g., trap intensity stabilization using active feedback; active force feedback; and the ability to generate multiple traps from a single laser via timesharing [44]. The trend in single-molecule instrumentation development is towards new methods that combine measurement capabilities, such as the recent combination of high-resolution optical tweezers with single molecule fluorescence measurements [50, 54]. This combination 38 was made possible by using an AO device to control a pair of traps via timesharing and interlacing with fluorescence excitation and detection [49]. Indeed, timeshared dual trap designs have the potential to be the most stable high-resolution optical tweezers (compared to, for example, polarization splitting methods [39]): the two trap beams share a complete spatial beam path and only experience an environment that differs in time on the timeshar- ing switching timescale (typically > 50 kHz), which is far faster than any optical component drift timescale. Figure 2.1: Interference patterns in acousto-optic device. Model of an acousto-optic device indicates the primary and reflected sound waves that produce the interference patterns responsible for the trap wiggles. However, AO devices have well known drawbacks: semi-periodic trap positioning and bead measurement inaccuracies commonly referred to as “wiggles”. Commonly used AO deflectors (AOD) have trapping inaccuracies on the 10 nm scale [41]. It has been speculated that these wiggles derive from sound wave interference patterns that vary sensitively with RF drive details. These inaccuracies are substantially reduced but not entirely removed by using an AO modulator, a simpler AO device compared to an AOD, but at the substantial cost of reduction in positioning range (approximately 10x reduced) and flexibility (e.g., 1D 39 instead of 2D positioning). Here we show that the AO device wiggles are indeed generated by the changing in- terference patterns of sound waves bouncing around in the AO device crystal as the trap positions are changed. We use direct digital synthesis (DDS) to produce the RF drive, and in addition to the typically controlled RF frequency and amplitude, we add to this direct digital control of the RF phase. We then periodically randomize the RF wave phase. The resulting decoherence removes the RF frequency dependent interference pattern and thus removes the trap positioning wiggles. We show that this method is effective for both AOM and AOD type devices, and indeed it should be applicable to all AO-based positioning methods. 2.2 Methods Two home-built optical tweezers instruments were used in these investigations: one con- taining an AOM (IntraAction ATM-803DA6B) and the other an AOD (IntraAction DTD- 274HD6C) to control the traps. The tweezers were both time-shared dual-optical traps constructed and operated generally as described in references [44] and [52], and shown in Fig. 2.2. An acousto-optic device rapidly deflects a trap laser between two positions to cre- ate two independent optical traps (each trap A and B is on for 5 µs duration, in alternating succession). The trap position and intensity are set by the RF frequency and amplitude driving the AO device. 40 Figure 2.2: Optical tweezers layout. Simplified layout of dual-trap optical tweezers setups used in this investigation containing an acousto-optic device. Orange and pink lines indicate trap laser and visible imaging illumination paths respectively. The trap laser passes through the AO deflecting device and then is expanded and imaged by a pair of telescopes (T1 and T2) onto the back aperture of the trap-forming objective lens (O1), which focuses the laser within a sample chamber (S). A small portion of the beam is picked off and sent to QPD2 for intensity monitoring and feedback. A second objective lens (O2) collects trap laser light exiting the sample chamber and passes it to a bead position detector (QPD1). Electrical control and measurement signals are indicated by black dashed lines. A FPGA chip is responsible for all measurement and control including the direct digital control of the DDS RF source. Using the established “coherent phase” method, the RF is synthesized via a direct digital synthesis (DDS) chip (Analog Devices AD9852 PCB Z) where RF parameters are digitally set for each trap on-cycle in real time by a field programmable gate array (FPGA) chip on a PC data acquisition and control card (National Instruments PCIe-7852). The synthesized RF is phase continuous (i.e., there are no discontinuities in the RF sine wave when changing the amplitude or frequency between traps). For the new “random phase” method, the RF phase is additionally set to a random value for each trap at the start of each on-cycle, synchronous with the change in RF frequency and amplitude (Fig. 2.3). 41 Figure 2.3: RF phase control. (a) The time-sharing dual traps were created by an acousto- optic device driven by RF frequency and amplitude. (b) The RF phase is continuous in the “coherent phase” method. The RF phase is set to a random value at the start of each trap cycle in the “random phase” method. The trap position is calibrated in the standard way by scanning a trapped bead, imaging the bead position and performing a linear fit to the result. Bead positions were measured once per trap on-cycle using the trapping laser itself via the back-focal-plane interferometry method [45]. Bead position measurement calibration is performed in the standard way via fitting to untethered bead Brownian motion power spectra [58]. While the AOD-deflected beam does not have a perfect Gaussian shape (see Fig. 2.8a), a bead in the AOD-produced trap does behave as if bound in a simple harmonic potential with a well-defined center, for sufficiently small excursions from the center (which is generally true for a trap produced by any method). 1 µm diameter polystyrene beads were used. Tethers consisted of 3 kb double stranded 42 DNA and were prepared using standard polymerase chain reaction (PCR) methods. 2.3 Results 2.3.1 AO device positioning inaccuracies Figure 2.4 demonstrates typical AO device positioning inaccuracies when scanning trap positions. Figure 2.4 shows simple force vs extension measurements of a pair of beads tethered together by a single 3 kb DNA molecule acquired by the AOM-based instrument. Curve (1) is obtained and plotted using standard methods. A very clear and substantial “wiggling” error is observed on what should be a smooth curve. This error can be mitigated to some extent by first acquiring an untethered bead position vs trap position measurement (e.g., the curves in Fig. 2.6) and subtracting that “offset curve” from the raw bead position measurements prior to position and force calibration. Then, as seen in Fig. 2.4 curve (2), the error wiggles mostly vanish. The offset curve must be acquired prior to tethering the beads, and there will be a delay before performing measurements. However, wiggle inaccu- racies are not stable and shift over time (Fig. 2.5) leading to inconsistent subtraction and the strong return of wiggles eventually (Fig. 2.4 curve (3)), thus introducing errors and limiting measurement time. 43 Figure 2.4: Force versus extension measurements pulling on DNA. Traces (1)-(3) used coherent phase methods: (1) “raw” data, (2) raw data with offset scan subtracted, (3) offset scan subtracted but after 30-minute delay. Trace (4) used random phase method. (Inset) difference between force and a simple polynomial model (error) for corresponding traces (1) (blue) and (4) (purple) showing zoom of wiggles and measurement resolution. Figure 2.5: Wiggles shift over time with acousto-optic device temperature. Bead measurement wiggle frequency drift (upper) and AOM device temperature versus time while trapping. Time t=0 is when trapping experiments started and the acousto-optic device was first powered on. 44 2.3.2 Untethered beads The untethered bead offset curves are shown in Fig. 2.6. Here a pair of untethered beads were trapped then their separation increased by moving trap A at a constant velocity while trap B remains stationary (as in, e.g., a force ramp measurement). Without a tether con- necting the beads, each bead should remain at the center of their respective trap and there should be no measured displacement of the beads during the scan. However, using the usual “coherent phase” AO device method (colored traces), the measured bead positions oscillate during the scan, for both the moving and stationary traps, in both the direction of trap motion (x) and orthogonal to it (y). The amplitudes of these wiggles vary by trap (moving versus stationary) and dimension (x versus y) and are as large as 1 nm and 25 nm for the AOM- and AOD-based tweezers respectively. For the AOM, the wiggle pattern is dominated by a single sine wave component, as apparent in the simple bead position versus trap position pattern (Fig. 2.6a) and also the Fourier power spectra of the patterns (Fig. 2.6b) which are dominated by single peaks. For the AOD, the pattern is more complex and composed of multiple significant sine wave components as seen both directly (Fig. 2.6c) and in the power spectra (Fig. 2.6d). It is especially surprising, and also revealing, that the undesired wiggles are present in bead measurements not only for the moving trap A, whose RF drive signal changes as the trap is scanned, but also the stationary trap B, whose RF signal remains fixed. The stationary trap also shows the simplest wiggle pattern (e.g., trap By (purple curve) measurement in Fig. 2.6a), with a wiggle period corresponding to the trap A RF frequency being incremented by 190 kHz. A detailed consideration of how the RF is applied to the AOM to create the traps provides insight. Each trap is created in sequence by instantaneously 45 changing the RF frequency without introducing a discontinuity in the RF sine wave output (continuous phase accumulation). Hence the starting phase of each trap RF signal is the final phase of the previous trap. The time interval of each trap is fixed. As the frequency of trap A is scanned, its final phase oscillates, and hence the starting phase of the stationary trap also oscillates. For the AOM-based instrument, trap A begins scanning at 80 MHz which corresponds to 400 RF cycles during the 5 µs trap A on duration. The phase shift is back to zero when 401 RF cycles fit perfectly into the 5 µs duration, which corresponds to 80.2 MHz, or an increase of 200 kHz, nearly matching the observed stationary trap wiggle pattern. A similar correspondence is also seen for the AOD-based instrument; however, the patterns are more complicated. This implies that the wiggles originate from a phase- dependent combination of the drive RF from successive trap on intervals, i.e., trap A wiggles derive from the superposition of the present trap A sound wave with the previous trap A sound waves and similarly for trap B. This is consistent with previous hypotheses that sound waves reflecting within the AO device crystal can combine and interfere to create wiggle patterns (Fig. 2.1). We were able to completely obliterate the trap scanning wiggle interference patterns by reducing the coherence of the sound waves within the AO crystal. We introduced dis- continuities in the RF signal phase at the start of both trap A and B intervals by specifying a random digital phase value synchronous with switching the RF frequency between traps. The resulting wiggle-free bead offset measurements (black traces) are shown in Fig. 2.6a and 2.6c for both the AOM- and AOD-based instruments respectively. The corresponding power spectra of the curves also show a dramatic elimination of the spurious wiggle patterns (Fig. 2.6b and 2.6d). Not surprisingly, wiggle errors are also completely eliminated in force versus extension measurements of tethered beads (Fig. 2.4 curve (4)). 46 Figure 2.6: Demonstration of positioning and measurement inaccuracies and the corresponding spatial power spectra for dual-optical tweezers setups. (a-b) AOM- based optical tweezers. (c-d) AOD-based optical tweezers. (a) and (c) show the bead position measurement for each bead (A versus B) and dimension (x versus y) while trap B was held constant, and trap A was scanned. For bead and dimension, there is a pair of measurements: upper (colored) used the coherent phase method, and the lower (black) is corresponding mea- surement after switching to the random phase method. (b) and (d) show the corresponding spatial power spectra of the Ax (blue) and Ay (red) bead position versus trap position using the coherent versus random phase. 47 2.3.3 Tethered beads The random phase method did not result in any noticeable degradation in the performance of the optical tweezers. In particular, the random phase method did not introduce any extra measurement noise compared to the coherent phase method as seen in the fixed trap position measurement of tether extension vs time in Fig. 2.7a, and the corresponding power spectra comparing the two methods in Fig. 2.7b. The offset in extension upon changing to the random phase method is due to the coherent phase method sampling a particular wiggle phase pattern for the particular fixed trap locations. Figure 2.7: Bead measurement noise for coherent versus random phase methods. (a) Tether extension versus time for both traps held at fixed positions. Initially the coherent phase method is used and then abruptly switched to the random phase method (indicated by dashed line). (b) Corresponding power spectra of the coherent phase (blue) and random phase (red) portions of the data in (a) are indistinguishable. Not all AO device interference effects are removed by the random phase method. In particular, interference for the AOD is more complicated than for the AOM. This is likely because the AOD uses a phased array set of piezo oscillators compared to a single piezo oscillator for the AOM. In addition to the bead measurement errors seen when positioning traps, the AOD also imposes a strong inference pattern visible in the trap laser beam profile. Figure 2.8 shows the trap laser beam profile following AOD deflection using the coherent 48 or random phase method. Using either the coherent or random phase method, the AOD de- flected beam profile shows a strong interference pattern compared to the initially Gaussian beam profile. The usual coherent phase method beam profile shows additional interference patterns that transiently change as the beam is scanned. However, when using the random phase method, the beam profile is fixed while it is scanned and there is no longer fine in- terference structure variation while scanning (as pointed out in Fig. 2.8c and 2.8d). It is when the interference patterns change with trap position that they cause measurement and positioning errors. 49 Figure 2.8: Trap laser profile after AOD and the corresponding Fourier power spectra. (a) Trap laser profile after AOD using coherent phase methods. (b) Trap laser profile after AOD using random phase methods. Images have same scale, as indicated by white bar in (b) whose length is the full width at half maximum (FWHM) of the beam profile. (c) and (d) are corresponding Fourier power spectra of images in (a) and (b) (log scale, spectra image pixels summed into 8 x 8 pixel bins). Images have the same scale, as indicated by the white bar in (d) whose length is 1/FWHM of (b). Arrows in (c) highlight transient structure that appears during AOD beam scans only when using the coherent phase method. 50 2.4 Discussion We have shown that AO device bead measurement inaccuracies systematically vary with trap positioning and can be completely removed by reducing the coherence of the AO device RF drive signals. This is a hallmark of wave interference where the observation of interference patterns depends on sufficient coherence between interfering waves. It is unlikely that the interfering waves are the electronic RF signals themselves. The reflections at RF electronic signal path interfaces are suppressed by many orders of magnitude and should be negligible given the quality RF cabling and components that were used, and the particular patterns we observe depend on the particular AO device under test and its optical alignment. Further, we have shown that dephasing the currently on trap interval from its previous on interval (>5 µs previous) is sufficient to remove the wiggles. The sound waves within the AO device crystal are resident in the crystal on this time scale: the speed of sound within the crystal (TeO2) is ∼4 mm/µs, while the crystal size is ∼5 mm and absorption of the sound waves at crystal surfaces is imperfect. We also observe that the wiggle offset curves exhibit a phase shift that is strongly correlated with the AO device temperature (Fig. 2.5). The shift with temperature is consistent with the thermal expansion of the AO device crystal and the subsequently changing phase of reflected waves relative to the incoming waves. This is all consistent with the physical model that the wiggling trap positioning error is due to fine structure of sound waves reflecting and summing within the AO device crystal. The shifts in the wiggles over time with the changing AO device temperature should be derived from slow drift in the sound field interference patterns. If the size of the crystal changes on the order of the wavelength of the sound wave, only 50 µm, then changes in the interference patterns will occur. Changing the temperature of the crystal will lead to 51 thermal expansion or contraction which will change the size of the crystal. AO devices are generally very susceptible to overheating and temperature drift as the weak acousto-optic effect requires very large input RF power (∼2 W in our case) while the insulating crystals are on the ∼1 cm size scale. For the devices we used, we found that heating to a steady temperature takes tens of minutes and that the temperature is very dependent on the recent on/off duty history. Ways to reduce wiggle drift are: 1) never turn off power to the AO device and 2) improve the temperature control of the AO device. However, by simply using our random phase method, the wiggles are completely obliterated and none of these concerns matter any longer. We explored other protocols to introduce random phase jumps in the RF signal to reduce the wiggles. Generally, these alternate methods introduced more phase jumps at random times during the trap on-cycles. Using the synthesis method described above, ∼5 additional phase jumps could be added during a 5 µs trap on-cycle. However, these phase jumps during the trap on-cycle, as opposed to at the start, introduced significant kicks to the trapped beads and dramatically increased the trap measurement noise. Similarly, random phase jumps can be introduced by using a fancy RF signal generator (e.g., the Stanford Research Systems SG380) as an RF reference clock and using built-in phase noise features. Wiggles are removed, but unacceptable measurement noise is added. We noticed that for the AOM as opposed to the AOD, the magnitude of the wiggle error is significantly larger in the y compared to the x dimension. Again, we note that the details of the AO device interference patterns, e.g., its amplitude, vary from device to device and very likely depend on the details of the device construction. We note that comparing an AOM to an AOD, the AOM device is asymmetric in that there is a single crystal operating in the x-axis direction only, while the AOD has a pair of crystals operating orthogonally in 52 x and y. 2.5 Conclusion While AO device wiggle inaccuracies do vary qualitatively and in magnitude from device to device and in different instrument setups, their straightforward elimination using our method provides broad benefits, particularly for high resolution applications. Force versus extension scans can be performed for long durations without concern for drifting wiggles that cannot easily be subtracted later in post-processing. Active force feedback methods which rely on real-time active and precise trap positioning are improved. Measurement drift generally is reduced because shifting wiggle patterns convert to shifting measurement offsets even for fixed trap measurements. Further our method allows AODs to be used in high-resolution optical trapping experiments, increasing the usable scan range by nearly 10-fold and allowing two-dimensional positioning. The application of the random phase method should also be widely used as it can be integrated into existing optical tweezer setups as a straightforward software change. 53 Chapter 3 Combined high-resolution optical tweezers and multicolor single-molecule fluorescence† We present an instrument that combines high-resolution optical tweezers and multicolor confocal fluorescence spectroscopy along with automated single molecule assembly. This hybrid instrument provides the detection of sub-nanometer mechanical displacements via optical tweezers with single-fluorophore sensitivity via confocal microscope. Three fluores- cence excitation lasers provide a wide spectrum range that covers the most commonly used single-molecule fluorophores. Multicolor allows the simultaneous observation of multiple proteins or F¨orster resonance energy transfer (FRET) measurements within a molecular complex. Simultaneous tweezers and fluorescence measurement is particularly challenging due to fluorophore photobleaching, even more so if multiple fluorophores are to be measured. We present the general design principles to overcome the challenges. We demonstrate the ability to simultaneously measure sub-nanometer extension changes and single fluorophore signals. We investigate the photostability of quantum dots and fluorophores. Lastly, we present the computer-controlled single molecule assembly line that enables the precise as- sembly of multi-molecule complexes while preserving fluorophores. †The work in this chapter will be published as: Cho-Ying Chuang, M. Zammit, M. L. Whitmore, J. L. West, and M. J. Comstock. Combined high-resolution optical tweezers and multicolor single-molecule fluorescence with automated single molecule assembly line 54 3.1 Introduction Single-molecule techniques have become powerful tools to investigate fundamental biological systems over the last two decades. In contrast to ensemble assays which provide measure- ments of the average behavior of molecules, single-molecule measurements provide another level of detail which can only be accessed by directly observing the behavior of individual molecules. Two major single-molecule methods approaches are force- and fluorescence-based measurements. Force-based measurements, as made by optical tweezers, magnetic tweezers or atomic force microscopes, can be used to observe the motions and conformations of indi- vidual proteins for long time durations over large distances with sub-nm resolution [29–31]. In addition, applied force can tune molecular stability and probe free energy landscapes. Single-molecule fluorescence methods can also be used to detect the motion and conforma- tion of individual molecules [46, 59]. However, biological molecules generally possess many important degrees of freedom and assemble into complexes composed of many interacting molecules. Any individual single-molecule method is generally limited in its view of this complexity. Recent progress has been made by devising instruments and methods that com- bine force and fluorescence measurements allowing the simultaneous observation of complex multi-protein or multi-degree of freedom systems [47,48,50,60–66]. A hybrid instrument that combined high-resolution optical tweezers with a single-color confocal microscope showed the promise of the method by resolving controversy and directly revealing the relationship of he- licase conformation and oligomeric coupled to DNA unwinding [50, 52]. While the helicase experiment above revealed unprecedented detail, its single excita- tion laser limited it to observing either a single protein conformation or counting a single protein species. To make progress, we need to increase the number of excitation and detection 55 channels integrated into the instrument. While multi-color confocal fluorescence microscopes have existed for some time, it is not trivial to integrate this into a high-resolution optical tweezers instrument. This is in part because 1) high powered optical trapping lasers tend to quickly photobleach fluorophores before measurement can take place and 2) high-resolution tweezers are complicated instruments that must be modified carefully in order to preserve their performance. Further, increasing the number of fluorophores used in a measurement increases the overall chance that one of them will bleach and end the measurement. Data throughput is crucial as high-resolution force measurements cannot be made in parallel. Here we present a multi-color hybrid instrument that combines three fluorescence ex- citation and detection channels with dual-trap high-resolution optical tweezers and enhanced computer automation. We precisely aligned the confocal spot by imaging inorganic fluores- cent quantum dots via a novel “cryptic binding” protocol. We evaluated the performance of the most commonly used organic fluorophores and surprisingly found that Cy5 and related fluorophores outperformed Cy3. We demonstrated the instrument performance by measur- ing the binding/melting of fluorophore-labeled short ssDNA probes to a tethered strand of DNA. To minimize photobleaching and improve the data acquisition efficiency, we devised a precise computer-controlled single-molecule assembly line. Trapping takes place within a sample chamber. Multiple adjacent laminar flow fluid streams contain uniquely labeled molecules which bind to a DNA tether. The computer automatically turns on the appro- priate excitation laser then quickly turn it off again upon molecule binding, and thereby moves the tether through the chamber precisely. The molecules were adding one by one while minimally wasting fluorescence and enabling maximal fluorophore lifetime. We believe these methods will be broadly applicable to many research groups investigating a wide range of protein-nucleic acid processing reactions. 56 3.2 Material and methods 3.2.1 DNA tether construct preparation Three types of DNA construct were used here: the “cryptic binding” hairpin construct, the “hybridization” construct, and the “fixed probe” construct. The cryptic binding hairpin con- struct was used for fluorescence scan imaging, the hybridization construct was used for DNA probe binding experiments, and the fixed probe construct was used for fluorophore bleaching lifetime experiments. These constructs were designed to have two similar double-stranded DNA (dsDNA) handles that have biotin and digoxygenin on opposite ends and a unique in- sert in the middle for particular experiments. Both dsDNA handles were produced by PCR reaction. The left handle “LH” was functionalized with biotin on the 5’ end, and the right handle “RH” was functionalized with digoxigenin on the 5’ end. The PCR products were then digested by the restriction enzymes to expose single-stranded DNA (ssDNA) overhangs. The sequence of the “Insert” was extended to hybrid to a complementary sequence on DNA handle. We annealed and ligated all segments together using T4 ligase, ran the ligation product on a 1% agarose gel, then purified and extracted the gel slices using a nucleic acid gel extraction kit (Qiagen). The concentration of the final product was determined using a Nanodrop UV-Vis spectrophotometer (Thermo Fisher Scientific). The digoxigenin and bi- otin ends of the DNA construct were linked to the anti-digoxigenin- and streptavidin-coated beads before the single-molecule measurements. The cryptic binding hairpin construct contained three DNA segments: two 1.5 kb dsDNA handles (LH and RH) and one 89-bp dsDNA hairpin insert (Insert). The final 3 kb 57 product has an 89-bp dsDNA hairpin in the middle. The hybridization construct contained three DNA segments: 1.5 kb dsDNA handle (LH), 19- to 47-nt ssDNA insert (Insert), and 1.7 kb dsDNA handle (RH). The final 3.2 kb product has an exposed ssDNA section in the middle. The fixed probe construct contained four DNA segments: 1.5 kb dsDNA handle (LH), 14- or 32-nt ssDNA insert (Insert), 3’ fluorophore-labeled complementary probe strands (Probe), and 1.7 kb dsDNA handle (RH). The final 3.2 kb product has a fluorophore-labeled probe in the middle. The “Probe” strands for both fixed probe construct and hybridization construct are the same, but the sequences of the “Insert” are slightly different. The “Insert” has 5-nt poly(dT) spacer on one side for the fixed probe construct, and 5-nt poly(dT) on each side for the hybridization construct. The shortened insert for the fixed probe construct allows us to anneal and ligate the phosphorylated fluorophore-labeled probes directly to the construct. This is for fluorophore property measurements so that we don’t have to worry about probes unbinding during measurement and giving a false bleaching signal. The sequences of the probe strands and inserts are shown in Appendix A. 3.2.2 Experiment conditions The DNA constructs were incubated with streptavidin-coated polystyrene beads (Spherotech Inc) for 30 minutes at room temperature for the biotin-streptavidin bonds to form, then di- luted in the sample buffer before single-molecule experiments. The single construct molecule was tethered in situ between a trapped streptavidin-coated polystyrene bead and a trapped anti-digoxigenin-coated polystyrene bead. Both beads were approximately 800 nm in diam- 58 eter. For the cryptic binding experiments, the sample buffer contained 100 mM Tris-HCl pH8, 50 mM NaCl, 20 mM MgCl2, 2 mM Trolox, 7.5 nM Qdot705-labeled probes (Qdot 705 streptavidin conjugate, Q10163MP, Invitrogen) and an oxygen scavenging system. For the fluorophore bleaching lifetime experiments, the sample buffer contained 100 mM Tris-HCl pH8, 50 mM NaCl, 2 mM Trolox, and an oxygen scavenging system. For the DNA probe binding experiments, the sample buffer contained 20 mM Tris-HCl pH8, 100 mM NaCl, 20 mM MgCl2, 2 mM Trolox, 5-10 nM fluorophore-labeled probes and an oxygen scavenging system. The oxygen scavenging system consisted of 1% glucose and 1% pyranose oxidase (or glucose oxidase) with catalase, and was used to improve fluorophore stability and tether lifetime [67, 68]. Trolox is a triplet-state quencher that reduces blinking and photobleaching [69]. Buffer conditions were chosen to optimize the probe strand binding rate. The single-molecule measurements were conducted at 23oC in a laminar flow cham- ber [70]. The custom-made sample chambers consisted of two No.1 microscope cover glasses (Fisher Scientific) sandwiching a sheet of Parafilm with channels cut out via a CO2 laser engraver. Experiments were conducted in a central channel flanked by a pair of side channels containing beads and connected by glass capillary tubes. The multi-color assembly line ex- periments were done in a five-channel chamber. The top channel contained anti-digoxigenin beads and the bottom channel contained DNA-construct-coated streptavidin beads. The central channel contained three adjacent flow streams. The three central channels merged smoothly to form a laminar flow. Three fluid flows containing different types of probes were injected by a computer-controlled syringe pump (Harvard Apparatus) pushing simultane- ously on three gas-tight glass syringes (Hamilton) to ensure smoothly flowing buffer streams that do not mix. 59 3.2.3 Data analysis Data analysis was performed using standard methods implemented via custom codes pro- grammed in MATLAB version 2016b (Mathworks). Data were recorded at 66.7 kHz and then boxcar-averaged to a final lower rate (20 ms per data point) for analysis and plot- ting. Standard polymer models of DNA [8, 71] were used to calculate the extension of DNA. The extension of dsDNA handles was computed by extensible worm-like chain model, and the extension of ssDNA insert was computed by extensible freely-jointed chain model. For dsDNA, the persistence length was 53 nm, the contour length per base pair was 0.34 nm, and the stretch modulus was 1200 pN. For ssDNA, the persistence length was 0.75 nm, the contour length per nucleotide was 0.59 nm, and the stretch modulus was 800 pN. For the single-molecule fluorescence measurements, the lateral set position of confocal spot was de- termined from the extension of the tether at a given force and the radii of the trapped beads (provided by bead manufacturer). In the fluorophore bleaching lifetime experiments, the measured fluorescence intensity and lifetime of each time trace were determined by derivative-based step finding algorithm and visually verified. The fluorophore lifetime and total photon number were determined from the mean of individual molecule distribution. Both distributions were well-described by single exponential models. Based on the maximum likelihood estimator, the likelihood of the data was the mean. Error bars were standard error of the mean. In the DNA probe binding experiments, the changes in tether extension were found at corresponding fluores- cence locations. 60 3.3 Instrument Design The special about this home-built hybrid instrument is that it prevents the combination- specific drawbacks that would sacrifice the sensitivity of either one, as well as taking ad- vantages of both force and fluorescence techniques. We used the dual-trap design optical tweezers [39], which are formed from the same laser beam, to achieve high-resolution force and displacement measurements. The instrument was located in a sound-proof, temperature- controlled, dark room and controlled remotely outside the instrument room to minimize the environmental noises. Since the dual traps are away from the sample chamber surface, we chose confocal microscopy rather than total internal reflection microscopy (TIRF) for the single-molecule fluorescence detection. The biggest challenge for combining optical tweezers and fluorescence is the fluorophore photobleaching. Our approach is to interlace fluorescence excitation and detection with time-shared dual optical traps to avoid enhanced photobleach- ing. Since interlacing trap and fluorescence excitation light sources is a critical feature, precise timing alignment and data acquisition is essential. A field programmable gate array (FPGA)-based PC card (National Instruments PCIe-7852R) was used for instrument control and data acquisition. The FPGA provides customizable software flexibility with precise timing control at 25 ns (40 MHz) resolution which is very important for the high-speed (µs-level) synchronous control. Custom instrument control software was programmed in LabVIEW version 2012 (National Instruments). The instrument consists of three modules: optical tweezers, three-color fluorescence confocal microscope, and bright-field imaging system. The schematic diagram of the in- strument is illustrated in Fig. 3.1. Each module has its own light sources and detectors. 61 A detailed parts list is provided in Appendix B. Please note that the detailed setup for optical tweezers is described extensively elsewhere [40,72]. Here, we focus on the design, con- struction, and demonstration of the multi-color fluorescence excitation/emission with optical tweezers. Figure 3.1: Schematic of the instrument layout. Multi-color optical tweezers instru- ment contains three modules: Optical tweezers (orange solid lines), three-color confocal fluorescence excitation and emission (green, red, and blue solid lines), and bright-field imag- ing system (magenta solid lines). The fluorescence excitation laser sources are on the second optical table. The coupled laser beams (473 nm, 532 nm, and 633 nm) are sent to the isolation room through a single-mode fiber. The trapping laser and fluorescence excitation (532 nm and 633 nm) lasers are interlaced by AOM1, AOM2, and AOM3, respectively. The fluorescence excitation 473 nm laser has built-in modulation mode. The asterisks indicate the conjugate image planes. The dashed lines indicate the back-focal planes of the objec- tives. The arrows indicate adjustable translational or rotational stages. A detailed parts list is available in Appendix B. 62 3.3.1 Optical tweezers A pair of traps was generated by time-sharing a 5W single-mode, polarization-maintaining 1064 nm fiber laser via an acousto-optic modulator (AOM). We note that the acousto-optic deflectors (AODs) are more commonly used to deflect the laser beam in optical trapping and have a larger defection range than AOM, but the trap positioning inaccuracies have limited the AOD application in high-resolution optical tweezers. By using AOM, these inaccuracies can be reduced but not completely removed. With the method we devised in Chapter 2 [55], the semi-periodic trap positioning inaccuracies are completely eliminated in both AOM- based and AOD-based optical tweezers. AOM1 was used to deflect the laser and thus position the traps by controlling the first-order diffraction angle via the radio frequency input signal (RF, 70-90 MHz). The trap laser intensity was monitored by QPD2 and was modulated via a proportional-integral- derivative (PID) feedback loop using AOM1. AOM1 was also used to interlace the traps with the fluorescence excitation by turning the trap beam ON and OFF out of phase with the excitation laser switching. Since the RF signal was digitally generated, when the RF signal was off, the traps were completely OFF. Only the undiffracted zeroth order beam emerged from the AOM. Then the zeroth order beam was captured by a beam dump. In order to maintain sufficient trap stiffness, we modulated the trap and fluorescent lasers at an overall rate of 66.7 kHz. For each 15 µs (66.7 kHz) duration cycle, two optical traps were created in sequence via time-sharing for the first and second thirds of the cycle, and then both were off for the final 1/3 of the cycle (Fig. 3.2). A high numerical aperture water immersion objective (O1) focused the trap laser light to create the traps within a fluid flow sample chamber. The trap laser light scattered by the 63 trapped beads was then collected by the back objective (O2) and sent to an IR-enhanced quadrant photodiode detector (QPD1) for bead position and force measurement. The posi- tions of the trapped beads were measured via the standard back-focal-plane interferometry technique. The signals were calibrated via the standard method of bead Brownian motion power spectra [58]. Figure 3.2: Interlacing and time-sharing of optical traps and fluorescence exci- tation lasers. In one interlacing cycle, we interlace at 66 kHz (15 µs). Two optical traps are on in sequence during first two time intervals and off during the third interval. The fluorescence excitation lasers are only on during the third time interval while the trap laser is off. Signals were measured by the photodetectors (feedback QPD and PDs) and recorded by a digital oscilloscope. 64 3.3.2 Multicolor single-molecule fluorescence A critical component of integrating the fluorescence microscopy to the optical tweezers is to co-align the fluorescence excitation beams with the trapping beam. The high-resolution optical tweezers alignment is sensitive. To simplify the coalignment of trap laser and fluo- rescence excitation lasers, we coupled three fluorescence excitation lasers into a single-mode fiber. The fiber has the property that all light coming out will be coaligned automatically. In this way, there is only one beam alignment with the trap. Fiber coupling not only simplifies the alignment of the multi-color confocal microscope, but also gives rooms for future exten- sion such as adding more excitation sources without realigning current setup. It also allows us to move a lot of fluorescence hardware off the main trap optical table, even to outside the room. This saves space on the main table and reduces noise/heat in the instrument isolation room. The downside is that the useable excitation laser intensity is lower because the input coupling efficiency cannot be 100%. In our case, we got 50% coupling efficiency. Generally, it is not a problem because usually a few µW excitation power is enough for single-molecule confocal fluorescence experiments. We always need to use neutral-density filters (F8, motorized flip filter) to reduce the laser intensity to desired power. Three fluorescence excitation lasers provide a wide spectrum range and clear spectral separation from blue to red: blue (473 nm), green (532 nm), and red (633 nm). It is designed to work with Alexa 488, Cy3, and Cy5 fluorophores or other substitute fluorophores, the most commonly used set of in vitro single molecule fluorophores (Fig. 3.3). We used visible achromatic doublet lenses in the multi-color confocal path to ensure consistent focusing and handling of all lasers and fluorescence across the entire visible spectrum. The fluorescence excitation lasers were interlaced with the traps. For the green and 65 Figure 3.3: Excitation and emission spectra of the fluorescent molecules. The excitation (dashed lines) and emission (solid lines) of Alexa 488 (blue), Cy3 (green), Alexa 647 (purple), Cy5 (red), and Qdot705 (orange). The five fluorescent molecules we studied in this Chapter cover the visible spectrum. red lasers this was done by modulating AOM2 and AOM3 respectively. The blue laser was a diode laser capable of fast intensity modulation by design. It was directly modulated with a digital input signal from the FPGA. Fluorescence lasers were only ON when both traps were OFF to minimize photobleaching [49]. Figure 3.2 shows a direct simultaneously measurement of the excitation and trap laser intensities during interlacing. Unlike to AOM1 for optical traps, the AOMs for fluorescence only controlled the intensity. The RF frequency did not change. Three fluorescence excitation lasers were coupled into a single single-mode polariza- tion maintaining optical fiber. Three excitation beams went through the same optical path, focused by the front objective (O1), and coaligned with the traps. The fluorescence excita- tion beams focused in between the traps in the same focal plane to excite molecules tethered 66 between the trapped beads. The emission fluorescence signals within the confocal volume were collected by the same objective (O1), then separated by dichroic mirrors, filtered by op- tical filters, and detected by three single-photon counting fluorescence detectors (avalanche photodiodes, APD). The timing synchronization (time alignment) between light sources and data acquisition is important. The APD only recorded the signals during the fluorescence excitation lasers ON period. Instead of using the APD gate digital input, we kept APD ON and programed the FPGA to record photon counts during the fluorescence period (i.e. the final 1/3 of the interlacing cycle). PD1, PD2, and PD3 are silicon photodetectors for monitoring fluorescence excitation laser intensity of blue, green, and red lasers respectively. The signals were sent to the FPGA and a proportional-integral-derivative (PID) feedback loop was used to stabilize the laser intensity. We used multiple methods to precisely co-align and position the fluorescence excita- tion beams relative to the optical traps both axially (z, the focal plane) and laterally (x and y within the focal plane). Coarse co-alignment of the excitation and trap lasers was achieved by observing focused trap and excitation laser spots on the visible microscope CCD camera. Axially positioning (focal depth) of the excitation laser was finely adjusted by translating the lens L3 of the telescope T3 along the beam path using a precision translation stage. Fine lateral positioning of the excitation lasers was controlled by a closed-loop piezo mirror stage (PM). Precise alignment was achieved and verified using complimentary methods. First, the excitation lasers can serve as back-focal-plane bead detection lasers (same as the trap laser). The piezo mirror raster-scans the excitation lasers across the trapped bead region while a position sensitive detector (PSD) records the detection image of a pair of trapped beads. When the lasers are co-aligned with the bead plane, the detection im- 67 ages of the beads are minimum diameter and circular without distortion (Fig. 3.4a). As expected, since three fluorescence excitation beams came out from fiber were coaligned and we used visible achromatic doublet lenses, it is the optimal position for all three fluorescence excitation beams (Fig. 3.4b). Next, we fixed the confocal lateral location at the center of one trapped bead and obtained the bead Brownian motion power spectra. The sensitivity of the measurement is maximized when the excitation laser is co-aligned in the same plane as the trapped bead. The PSD voltage to position conversion factor α (nm/V) in the lateral x and y directions were obtained by fitting the power spectrum in the standard way (just as is done for the optical trap calibration). If we obtain conversion factor α versus focal depth, the optimal position coincides with the minimum of α (i.e. the highest detection sensitivity). Figure 3.5a shows the calibration for the same bead using trap beam (left) and fluorescence beam (right) as the detection laser. The chamber was aligned relative to the trap beam, so there was no low frequency noise on both x and y direction using trap detection beam. The low frequency noise for the fluorescence detection beam may be from the slightly tilt angle between trap beam and fluorescence beams. Finally, the focal depth alignment was confirmed with the APD using fluorescence emission signals. The image of a trapped fluorescence bead was scanned using the piezo mir- ror stage (Fig. 3.5b). The optimal position would have the maximum fluorescence intensity. The axial alignment by both PSD and APD methods agreed with each other (Fig. 3.5c). 68 Figure 3.4: PSD scan images from all three fluorescence excitation lasers using the piezo mirror stage. (a) Scan the confocal spot over one trapped bead. The deflection in x (upper) and in y (lower) of the green laser when it scans. Red represents positive deflection and blue represents negative deflection. (b) Scan the confocal spot over two trapped beads. The deflection in y direction when the blue (top), green (middle), and red (bottom) lasers scan. 69 Figure 3.5: Axial alignment of confocal spot. (a) Bead calibration with trap detection laser (left) and fluorescence detection laser (right) in x (upper) and y (lower) direction at the optimal focal depth position. The power spectra are from the same trapped bead. (b) Scan images of a trapped fluorescent bead using piezo mirror stage at the optimal focal depth position. The emission spectrum of the fluorescent bead (F8819, Invitrogen) covers both green and red APDs detection wavelength. (c) Adjustment of confocal spot focal depth. (Upper) The conversion factors αx (blue circles) and αy(red crosses) were determined from the power spectrum fitting in (a). (Lower) The fluorescence signals from green APD (green circles) and red APD (red crosses) were obtained from the APD scan images in (b). 70 3.3.3 Bright-field imaging The bright-field imaging system consists of an infrared 940 nm LED as the light source and an IR-enhanced CCD camera to visualize the sample chamber, beads, trap manipulation, and tether formation. F1 (short pass, 1000 nm) and F2 (long-pass, 850 nm) were used to remove excess trap and fluorescence laser intensity as needed. 3.4 Results and discussion 3.4.1 Cryptic quantum dot binding for robust signal confocal align- ment Precise lateral alignment of the confocal spot with the tether construct is essential for efficient single-molecule fluorescence detection. To confirm the lateral alignment with a tethered construct, the final way is imaging a single fluorescence molecule in between the trapped beads. While it is possible to directly image a single organic fluorophore such as Cy3 in our setup and thereby precisely position the confocal laterally [50], it is extremely challenging to obtain images by raster scanning rather than with a camera due to the practical limitations of photobleaching. Inorganic fluorescent quantum dots are semiconductor particles that have brighter emission signals and less likely to photobleaching compared to the organic fluorophores [73]. Also, quantum dots have a wide excitation spectrum and a narrow emission peak while the organic fluorophores have a narrow excitation spectrum and a narrow emission range. We took these advantages and developed a new fluorescence detection scheme that is useful for 71 both instrument alignment and measurement. We used Qdot 705 (emission maximum ∼705 nm) in our demonstration experiment. Qdot 705 can be excited by our three fluorescence excitation lasers and detected by the red emission detector. Using replenishable fluorescent-labeled probes in solution allows fresh probes to re- bind to the tether after photobleaching. However, a major challenge for diffusing fluores- cent probe experiments is the signal-to-background-noise ratio. Some of this background comes from probe bound to excess untethered DNA attached to the trapped streptavidin bead. To remove the fluorescence background from beads, we combined a multi-channel flow chamber and internal hairpin binding construct to perform “cryptic binding” and “dipping” experiments (Fig. 3.6). The hairpin constructs reduce fluorescence background from DNA- construct-coated beads because the only complementary binding site for the probes is the opened hairpin tether. Once the probe binds, the tether extension changes are observed by the optical tweezers. We simultaneously measured fluorescence and extension for a hairpin tether (Fig. 3.7). When hairpin tether opened, it exposed a single-stranded portion near the center to the complementary quantum-dot-labeled probe. The binding of quantum-dot- labeled probe changed the tether extension and increased the fluorescence signals. 72 “Cryptic binding” experiments with quantum dots. Hairpin con- Figure 3.6: struct creates “cryptic binding” site to reduce fluorescence background from beads. Two polystyrene microspheres (gray spheres) are trapped by dual optical traps (orange cones) and tether together by the DNA construct. When the complementary quantum dot-labeled probes bind to the tether, the fluorescence signals (red disk) are detected by the confocal microscope (blue cone) and the tether extension changes are monitor by optical tweezers. 73 Next, we performed dipping experiments to image inorganic fluorescent quantum dots (Fig. 3.8). The tether was moved to the “blank” buffer channel. The fluorescence excitation lasers were then turned on to collect the signals. The confocal scan image shows the accuracy of the calculated set spot that computed from the bead size and the DNA tether extension. The differences between the center of quantum dot and the set spot were only within 0.1 µm. The excellent brightness and photostability of quantum dot make it a good candidate for fluorescence alignment. Although quantum dots blinked a lot, it did not bleach after exciting high laser power 750 µW fluorescence laser for minutes (multiple scan images). Dramatically reduced photobleaching compared to organic fluorophores allows imaging and precise fluorophore localization. This “cryptic binding” and “dipping” method significantly reduced the background fluorescence signals from beads and buffer. This method can be applied to experiment need long fluorescence probe lifetime or high brightness. 74 Figure 3.7: Representative time trace of hairpin cryptic binding experiments. Fluorescence excitation laser turned on around 5 seconds. Hairpin construct opened around 15 seconds and revealed a ssDNA section in the middle. Quantum dot didnt bind until 66 seconds. (Insets) A zoomed view of the force (upper panel), tether extension (middle panel), and fluorescence signal (lower panel) changes when quantum dot bound to the exposed ssDNA portion. 75 Figure 3.8: Confocal alignment scan images using quantum dots. (Upper) The confocal scan image from PSD. The fluorescence beam deflected by two trapped beads as it scanned across the beads. The red circles indicate the size of trapped beads, the red dashed line represents the DNA tether, the blue cross is the bound probe position calculated from the polymer model, and the white crosses are the center of quantum dot determined from multiple APD scan images. (Lower) the confocal scan image from APD. The bound quantum dot probe stayed bright and survived for minutes after multiple raster scans. 76 3.4.2 Multicolor DNA probe binding demonstration In order to demonstrate the new instrument capabilities, we measured the binding and unbinding of fluorophore-labeled single-stranded DNA oligonucleotides to a strand of DNA tethered between a pair of trapped beads (Fig. 3.9). The tethered DNA had a middle section with an exposed ssDNA probe binding site. Short complimentary ssDNA probe molecules labeled with a variety of fluorophores were dispersed in the surrounding solution. Upon binding of a single probe molecule to the tethered DNA, the ssDNA section of the tether is converted to dsDNA, which has a different extension generally (Fig. 3.10), and thus is detected by the optical portion of the instrument. At the same time, the confocal fluorescence portion of the instrument detects the presence of the fluorophore. Figure 3.9: Multi-color single fluorophore-labeled oligonucleotide hybridization experiment. When the complementary fluorophore-labeled probes bind to the tether, the fluorescence signals (red disk for red laser excited signals) are detected by the confocal microscope (green, red, and blue cones) and the tether extension changes are monitor by optical tweezers. 77 Figure 3.10: Force dependent DNA extension change upon hybridization. Force dependent extension change from 9-nt ssDNA to 9-bp dsDNA was calculated using extensible worm-like chain polymer model. The end-to-end extension is shorter (longer) when the force is above (below) the critical force 5.5 pN. Figure 3.11a shows an example simultaneous tether extension and fluorescence measurement for a Cy3 labeled probe excited by the corresponding green laser. A de- crease/increase in tether extension is observed simultaneously with an increase/decrease in the fluorescence signal upon probe molecule binding/unbinding from the DNA tether. In Fig. 3.11b, we plot the change in tether extension upon fluorophore binding and unbinding showing clear sub-nanometer trap measurement resolution similar to the previous instru- ment [50]. Example time traces (Fig. 3.12) show detection of red laser excited Cy5-labeled probes, green laser excited FRET labeled probes. The blue laser excited Alexa Fluor 488 probes are also shown. Note that consistent with other fluorescence-based instruments, blue laser excited fluorophores while measurable and usable provide substantially worse perfor- mance than red and green. 78 Figure 3.11: Simultaneous measurements of fluorescence and tether extension. (a) (Upper) The tether extension difference between single-stranded and double-stranded DNA. (Lower) Cy3 fluorophore labeled probes were excited by green laser. The binding and unbinding events were simultaneously observed by the stepwise increase and decrease of the fluorescence signals and the changes in tether extension. (b) Histogram for many binding and unbinding events of the recorded extension differences from the molecule in (a). 79 Figure 3.12: Multi-color single fluorophore-labeled oligonucleotide hybridization experiment under 10 pN force. (a) Cy3-Cy5 FRET labeled probes were excited by green laser. (b) Alexa Fluor 488 fluorophore labeled probes were excited by blue laser. (c) Cy3-Cy5 FRET labeled probes were excited by red and green lasers separately. At 0-80 seconds, Cy5 from the FRET probes were excited by the red laser. Red laser was off and green laser was turned on between 370-435 seconds. At 375-395 seconds, Cy5 signals were obtained through FRET. At 400 seconds, Cy5 bleached and only Cy3 signals were detected. 80 3.4.3 Cy5 outperformed other organic fluorophores We next quantitatively evaluated the performances of the most commonly used organic fluorophores for blue, green and red excitation in our instrument. We prepared tethers in the wet lab with the fluorophore-labeled probe DNA already bound and permanently ligated to the tether (Fig. 3.13a). This way the loss of fluorescence signal unambiguously indicates photobleaching rather than the unbinding of the probe (Fig. 3.13b). Experimental setup for fluorophore lifetime experiments. Figure 3.13: (a) Schematic of experimental setup. Utilizing fixed fluorophore construct to characterize fluo- rescence performance in high-resolution multicolor fluorescence optical tweezers. A variety of fluorophores were ligated to DNA construct. (b) Typical time trace of single fluorophore lifetime measurements. Laser was on at 5 sec, excited the fluorophore and given a 10 kHz signal. The fluorophore bleached at 23 sec. Total photons collected is the multiplication of intensity and photobleaching lifetime. We first investigated the effect of the trap on fluorescence performance using the most commonly used fluorophore, Cy3. The trap laser power was varied more than 10-fold over the typically used range. The corresponding trap stiffness was from 0.05 to 0.6 pN/nm while the power of 532 nm excitation laser was kept constant at 16 µW. As seen in Fig. 3.14a, the fluorophore emission intensity and lifetime was essentially unchanged by trap laser power, confirming that the interlacing method successfully protected the fluorophores from the trap laser. Cy3 can provide ∼45,000 photons before photobleaching. Next, we measured the effect of the fluorescence excitation laser intensity on Alexa Fluor 488, Cy3, Cy5, and Alexa Fluor 647 fluorophores (Fig. 3.14b). The fluorescence 81 intensity is proportional to the excitation laser intensity (Fig. 3.14b upper panel), but the fluorescence intensity at 32 µW were 10% lower than expected linear fit. Although we use the low power range in most experiments, this would be a concern at higher excitation power which might be used for high time resolution experiments. Bleaching rate which is the inver- sion of lifetime (Fig. 3.14b middle panel) is proportional to the excitation intensity. Both intensity and bleaching rate indicate that bleaching is mostly a single photon effect. There- fore, the higher the excitation laser intensity, the lower total photon numbers (Fig. 3.14b lower panel). Surprisingly, we found that the Cy5 performed much better than other fluo- rophores, particularly Cy3. Cy5 has the most photon numbers. It is tenfold (400k at 4 µW red laser) higher than Cy3 (80k at 4 µW green laser) and is comparable to the photo num- ber collected from other fluorescence-based single-molecule instrument such as total internal reflection fluorescence microscope (TIRF). Alexa Fluor 647 has similar intensity as Cy5, but slightly shorter lifetime. Alexa Fluor 488 has significantly shorter lifetime and fewer photon number (3k at 2 µW blue laser). Therefore, in order to optimize the performance, red exci- tation of Cy5 should be prioritized over green excitation of Cy3. 82 Figure 3.14: Fluorophore lifetime experiments with fixed probes. (a) Effect of the trap laser intensity on Cy3 photobleaching at 16 µW 532 nm excitation. n = 29, 25, 28, and 27 for trap stiffness 0.05, 0.2, 0.3, 0.6 pN/nm. The trap stiffness is proportional to trap power. (b) Effect of the power of fluorescence excitation lasers at 0.3 pN/nm trap stiffness. The blue-absorbing Alexa Fluor 488 (blue plus) was excited by 473 nm laser. n = 30 for 2 µW. The green-absorbing Cy3 (green diamond) was excited by 532 nm laser. n = 26, 14, 28, and 12 for 4, 8, 16, and 32 µW. The red-absorbing Cy5 (red circle) and Alexa Fluor 647 (pink cross) fluorophores were excited by 633 nm laser. n = 22, 28, 28, and 25 for 4, 8, 16, and 32 µW. n = 17, 23, 21, and 20 for 2, 4, 8, and 16 µW. Error bars are smaller than symbol sizes. 3.4.4 Automated single molecule assembly line Due to the finite photon number, it is critical to efficiently excite the molecule. We have combined multi-channel sample chambers with precise computer control of fluorescence mea- surement and triggered chamber motion to implement an automated molecular assembly line. This allows us to precisely add individual molecule of different types to a single DNA tether while conserving fluorescence photons and reducing photobleaching. We used Laminar flow chambers and put different fluorescently labeled molecules for assembly in different channels 83 (Fig. 3.15a and 3.15b). Each fluorophore-labeled molecular component is in a separate buffer and does not mix. The assembly is divided into two routines: detection and motion. Computer controlled excitation lasers on and off to detect the molecule binding events. The trapped complex can be moved between channels to add components or change buffer con- ditions. The software was programed to use the chamber motions as the trigger to turn the fluorescence excitation lasers on and use the fluorescence emission signals as the trigger to turn the fluorescence excitation lasers off. Figure 3.15: Multi-channel flow chamber. (a) Photograph of a 5-hole flow chamber. The top channel contains anti-digoxigenin beads. The bottom channel contains DNA-construct- coated streptavidin beads. The central channel consists of three laminar flow streams. (b) Laminar flow layout. The central channel contains three different buffers: buffer without probes (“Blank”), buffer with Cy3-labeled molecules (“Molecule 1”), buffer with Cy5-labeled molecules “Molecule 2”). Fig. 3.16 shows an example of our molecular assembly line. A single tether is formed in the “Blank” buffer without any fluorophores, then moves the tether to the next channel. The green laser is turned on automatically once the tether arrives the set spot. The tether is sitting in the buffer containing Cy3 labeled molecules and waits for the probes to bind. When the signal exceeds the pre-set background threshold level, it automatically turns the green laser off, moves to next channel, turns red excitation on, and repeats the process until we get all the components. We use computer program carefully turn laser on and off to 84 maximize the photon number and minimize the time of unnecessary exposure to excitation laser. In multi-fluorophore trapping experiments, multiple fluorophores need to be alive dur- ing measurement. Therefore, an automated computer-controlled assembly line provides fast, accurate, and efficient measurement than manual controlled measurement. Figure 3.16: Molecular assembly line with the multi-channel flow chamber. Ex- ample time trace for in situ molecular assembly. The DNA tether was formed in “Blank” buffer, then (1) moved to “Molecule 1” buffer, turned green laser on and waited for the probe binding, (2) turned green laser off once the probe bound, moved to “Molecule 2” buffer, (3) turned red laser on and waited for the probe binding, (4) turned green laser on and ob- serve FRET, (5) Cy5-labeled probe bleached or unbound, only Cy3-labeled probe showed emission, (6) Cy3-labeled probe bleached or unbound. No fluorophore emission. 3.5 Conclusion We have demonstrated a multi-color hybrid instrument combined high-resolution optical tweezers and multi-color confocal fluorescence microscope has a potential to be used to study the structure, interactions, and dynamics in real biological systems. We showed the robust and bright quantum dots are a great alternative fluorescent probe in the high-resolution multicolor fluorescence optical tweezers experiments and alignment. We have measured fluo- 85 rophore photobleaching lifetimes with a set of fluorophores for a variety of excitation powers in a range of buffer conditions. We also showed that Cy5 has a much better performance over other fluorophores. In addition, the throughput of the single-molecule measurements can be improved by our instrumentation development. The multicolor hybrid instrument and methods we developed open an opportunity to uncover the details of the complex biological systems. For instance, in addition to measuring the DNA tether extension changes (e.g., from helicase or polymerase motion), we can detect multiple fluorophore-labeled proteins in DNA replication which involves helicase, polymerase, and single strand binding protein. We can also study conformational changes in RNA polymerase during transcription via interfluo- rophore distance (FRET efficiency). These instrumentation advancements should enable the precise single-molecule assembly and measurement of complex, multi-component molecular machine systems. 86 Chapter 4 Mechanical folding pathway of the thrombin-binding aptamer G-quadruplex‡ § We performed high-resolution optical tweezers to directly observe folding and unfolding re- actions of thrombin-binding aptamer G-quadruplex (TBA-GQ) at very low forces (as low as 2 pN). Studying the kinetics of G-quadruplex formation is essential for gene regulation and therapeutic approaches for disease. TBA-GQ is the smallest G-quadruplex. It contains only two G-tetrads that make it less stable than other G-quadruplexes under force. Utilizing the sensitivity and stability of our optical tweezers, we were able to directly measure the folding and unfolding rate constants of TBA-GQ under force. We studied the folding and unfolding of TBA-GQ over a range of conditions including the impacts of force, monovalent salt ions, and thrombin. We also identified the transition state structure under a variable salt concentration. ‡The work in this chapter will be published as: Cho-Ying Chuang, H. Balci, and M. J. Comstock. Mechanical folding pathway of the thrombin-binding aptamer G-quadruplex by single-molecule force spec- troscopy §Some work in this chapter has been published as [74]: G. Mustafa, Cho-Ying Chuang, W. A. Roy, M. M. Farhath, N. Pokhrel, E. Anthony, M. J. Comstock, S. Basu, H. Balci. Biosensors and Bioelectronics, 121, 34-40 (2018) 87 4.1 Introduction G-quadruplex is a non-Watson-Crick secondary structure formed by single-stranded guanine- rich DNA or RNA sequences. Four guanines are held in a plane by Hoogsteen hydrogen bonds [75] and coordination with a monovalent metal ion such as N a+ or K+ to form a G-tetrad (Fig. 4.1a). Two or more G-tetrads can form in sequence and stack on top of one an- other, with the ion sandwiched in between, to form G-quadruplex structures. G-quadruplex structures are believed to be involved in several key biological processes such as replication, transcription, translation, and telomere maintenance [15, 76–79]. Of particular importance, telomeres, the end regions of eukaryote chromosomes, are rich in the G-repeat sequences known to form G-quadruplex structures in vitro. It was demonstrated that telomeric G- quadruplex protects the end of eukaryotic chromosomes and inhibits telomerase activity [80]. These structures are dynamic. In order to understand how they interact with proteins like telomerase, we should first characterize their folding and unfolding dynamics in isola- tion. Understanding the dynamics of G-quadruplex formation is crucial for both molecular biology understanding and structure-based drug design [81–84]. Thrombin-binding aptamer G-quadruplex (TBA-GQ) is a minimal model system of G-quadruplex. It is a 15-mer single-stranded DNA that folds into a chair-like two-layer G- quadruplex (Fig. 4.1b) [85, 86]. It is an important DNA structure in its own right as it specifically interacts with thrombin to inhibit blood coagulation [26] and has been investi- gated as a potential anticoagulant aptamer-based drug [25, 87]. TBA-GQ is one of the most studied G-quadruplex forming aptamers [88–99], however its folding/unfolding mechanism remains unresolved. Previous single molecule investigations have included probing the sta- bility of TBA-thrombin interactions and the role of salt in stabilizing structure [100, 101]. 88 Force spectroscopy can provide powerful methods for identifying folding and unfolding free energy landscapes and transition states, however TBA-GQs small size and relative instability under force (unfolding at < 5 pN) have thus far been limiting. Most recently the folding free energy landscape and transition states of TBA-GQ have been predicted via computational simulations [102–104]. These molecular dynamics (MD) simulations predict that G-hairpin and G-triplex are the possible intermediate states in TBA-GQ formation. Here we present a high-resolution optical tweezers investigation of TBA-GQ where we directly observed both folding and unfolding reactions at very low forces. Equilibrium condition measurements allowed us to directly and unambiguously determine folding and unfolding rate constants under varying conditions. Exploration of salt dependence showed a very strong increase in the folding rate constant with increasing salt while unfolding de- pendence was weak. Thrombin showed the opposite effect where increasing the thrombin concentration was seen to increase the fraction of TBA-GQ folded by decreasing the unfold- ing rate constant. The present of thrombin is stabilizing TBA-GQ while not changing the folding rate constant. Measurements of rate constants vs applied force probed the folding and unfolding free energy landscape. We were able to model both the folding and unfolding equilibrium data self-consistently using a single transition state clearly identified as the G- triplex structure. Altogether, these investigations present a clear picture of the folding and unfolding dynamics of TBA-GQ. 89 Figure 4.1: Structure of thrombin-binding aptamer G-quadruplex (TBA-GQ). (a) G-tetrad, the building block of G-quadruplex, contains four coplanar guanines that coordi- nate with a monovalent metal ion. (b) Schematic structure and dimensions of the chair-like two-layer thrombin-binding aptamer G-quadruplex. 90 4.2 Material and methods 4.2.1 Dual-trap high-resolution optical tweezers All measurements were performed using a home-built timeshared dual-trap high-resolution optical tweezers described in references [50, 105]. Two independent optical traps (66.7 kHz modulation rate) were created by rapidly deflecting a 1064 nm laser (IPG) via an acousto- optic modulator (AOM, IntraAction). A constant trap stiffness (0.2-0.3 pN/nm) was main- tained through a PID (proportional-integral-derivative) feedback loop using the same AOM to stabilize the trap laser intensity during trap positioning. Bead position and trap stiffness calibrations were performed via the standard analysis of bead Brownian motion power spec- tra [58]. Instrument control and data acquisition were performed using a field programmable gate array (FPGA)-based PC card system (National Instruments PCIe-7852R). Instrument control software was programmed in LabVIEW version 2012 (National Instruments). 4.2.2 Construct preparation The TBA-GQ insert contained 15 nucleotides (5’-GGTTGGTGTGGTTGG-3’) with 2-nt poly(dT) linkers on each side. Linkers are added to reduce the interaction of G-quadruplex with double-stranded DNA (dsDNA) handles. The insert was annealed and ligated to a biotin labeled 1.5 kb dsDNA handle on one side and a digoxigenin labeled 1.7 kb dsDNA handle on the other side. Both handles were synthesized by PCR reaction followed by re- striction digestion. The final product is a 3.2 kb dsDNA plus 19-nt single-stranded DNA (ssDNA) central region (Fig. 4.2). 91 Figure 4.2: Schematic of the TBA-GQ experimental setup using optical tweez- ers. The TBA-GQ construct (cyan) is tethered between two optically trapped beads (gray spheres) via two long dsDNA handles (red). 4.2.3 Experimental procedure A pair of trapped polystyrene beads were tethered in situ via the single construct molecule (Fig. 4.3 inset). For each tethered construct, we first performed non-equilibrium force ramp measurements at a scan speed of 100 nm/s and fit to the worm-like chain model for con- firming the proper behavior of the TBA-GQ. Three representative force-extension curves are shown in Fig. 4.3. The dashed lines represent polymer models of TBA-GQ folded (red dashed) and unfolded (black dashed). For good molecules, next, low force equilibrium mea- surements were obtained by holding the construct at fixed trap separations corresponding to desired mean forces. Data were collected at a set of trap positions to probe the folding and unfolding at set of mean forces. All data were collected at 23oC in a buffer containing 50 mM Tris-HCl pH 8, 150-1500 mM KCl, and 0-1200 mM NaCl. An enzymatic oxygen scavenging system (1% glucose and 1% glucose oxidase with catalase) was also included in the buffer to prevent tether damage. For thrombin experiments, the buffer contained 300 mM KCl, 150 mM NaCl, 50 mM Tris-HCl pH8, 0-150 nM human alpha-thrombin (Invitrogen RP-43100, 92 Thermo Fisher Scientific Inc.), and the enzymatic oxygen scavenging system. The sample channel consisted of two adjacent laminar flow streams at a constant flow rate 100 µL/hr. The tethers can be transferred between the adjacent streams (with and without thrombin) in less than 3 seconds. Figure 4.3: Single-molecule optical tweezers study of thrombin-binding aptamer G-quadruplex (TBA-GQ). Non-equilibrium force ramp measurements of TBA-GQ in the presence of 600 mM KCl. The inset cartoon shows a single TBA-GQ folded and unfolded under mechanical force via optical tweezers. Three representative force-extension curves for pulling (blue) and relaxing (red) are shown with horizontal shifts for clarity. The extension of DNA tether follows the polymer model of TBA-GQ folded (red dashed) and unfolded (black dashed). 93 4.2.4 Data analysis Tether polymer model Standard polymer models of DNA were used for force-extension curve fitting. dsDNA handles and ssDNA were modeled using the extensible WLC model [7, 8, 11, 106, 107]. For dsDNA, the persistence length was 53 nm, the contour length per base pair was 0.34 nm, and the stretch modulus was 1200 pN. For ssDNA, the persistence length was 1.1 nm, the contour length per nucleotide was 0.59 nm, and the stretch modulus was 800 pN. The ssDNA per- sistence length 1.1 nm was determined by empirical fit to non-equilibrium force ramp data and equilibrium time traces. These parameters fit our data well over a range of force and salt concentration. Rate constants and transition state fitting Data were analyzed using standard methods implemented via custom codes programmed in MATLAB version 2016b (Mathworks). Data were collected at 66.7 kHz then boxcar- averaged to a final lower rate (between 9 to 90 ms per data point) for analysis. Automatic step finding was performed by a statistics-based changepoint method and visually verified to prevent artificial events. The dwell time distributions were plotted as binned histograms and unbinned cumulative probability. Both were well-fit by single exponential models (Fig. 4.4). Therefore, rate constants were estimated by taking the inverse of the mean dwell time of the distribution, which is the maximum likelihood estimate. Error bars were standard error of the mean. 94 Figure 4.4: Folding and unfolding dwell times show single exponential distribu- tion. (a-b) Dwell time histogram of the (a) folded states (i.e. unfolding rate constants) and (b) unfolded states (i.e. folding rate constants) from one example time trace at 300 mM KCl. Since the data is single exponential distribution, the maximum likelihood estimate (MLE) of time constant τ is the mean of the dwell times. (c-d) Logarithm plots of the histogram with least squares linear fits. (c) is corresponding to (a), and (d) is corresponding to (b). The linear fit at log scale indicates that the data is single exponential distribution. (e-f) Cumulative probability distributions. The data (blue lines) show very good agreement with MLE (red lines) and 95% MLE confidence intervals (green dashed). 95 However, we do not directly fit to these rate constants when performing Bell fitting or modeling the transition state. Instead we implemented a generalized maximum likelihood global fitting method to analyze the equilibrium data and the corresponding force-dependent rate constant models. The maximum likelihood method is a powerful parameter estimation method for a given model and a set of data. It maximizes the value of the data by using each individual dwell time at its corresponding force rather than the simple mean. Different molecules were measured for different amount of time and have different number of events. For each dwell time, we summed the log of the probability of the measured dwell times. Then iteratively find the parameter values that best describe the data. The uncertainty of the fit parameters was estimated from the standard bootstrap resampling method. 4.3 Results 4.3.1 Force dependent In order to perform initial characterization of TBA-GQ unfolding and verify that we have good individual TBA-GQ tethers we performed standard force-ramp experiments where we repeatedly oscillated the trap separation to pull and relax the tethers. A set of typical force vs. extension curves are shown in Fig. 4.3. TBA-GQ tethers began in the folded state at low force and then unfolded with increasing force in single steps (a simultaneous increase in extension and decrease in force). For these particular conditions, 100 nm/s scan speed, refolding occurred at low force and was only sometimes observed. Intermediate states were not observed. The observed change in contour length before and after the observed ruptures closely matches the expected change in contour length for a single TBA-GQ molecule. Thus, 96 we verified that our tether constructs each contained a single TBA-GQ molecule. We then probed the TBA-GQ unfolding and refolding landscape by directly measuring these reactions under equilibrium conditions with fixed trap separation. While we could perform similar measurements via out-of-equilibrium force ramp measurements, as others have commonly done in the past, the data from such measurements have to be analyzed with careful consideration of the detailed force ramp conditions. High ramp rates and forces can distort free energy landscapes so much that completely different paths and states are observed as compared to zero force and equilibrium [108–110]. However, the high-resolution tweezers allow us to perform equilibrium, low force measurements of TBA-GQ. Figure 4.5a shows a typical equilibrium measurement of TBA-GQ. The trap separation remains fixed except for sudden increases that increase the average force (e.g., at t = 200 s, there is a stepwise increase in the trap separation and the mean force changes from 2 to 3 pN). At a given mean force, stepwise increases and decreases in the force signal correspond to TBA-GQ unfolding and refolding respectively. The reaction was dominated by two states, with no reproducible intermediates. The extensions of the two states agreed with the two states observed in the force ramp measurements and the polymer modeling of the expected fully folded and unfolded TBA-GQ contour lengths. Figure 4.5b shows the raw data distributions in Fig. 4.5a versus mean force. Two peaks are observed corresponding to the folded and unfolded TBA-GQ states. With increasing mean force, weight shifts from the folded to the unfolded peak. The equivalence force where TBA-GQ was equally likely to be found folded versus unfolded was approximately 3 pN (for high salt, 1000 mM KCl; we discuss deviations with varying salt below). The unfolding rate constants increased with increasing force and the folding rate constants decreased with increasing force as shown in Fig. 4.5c. A simple Bell model (black dashed) [111] of exponential dependence of rate 97 constants on force can provide a useful phenomenological description of the data. k = k0 exp( ∆x · F kBT ) (4.1) where k is the rate constant, k0 is the rate constant at zero force, ∆x is the distance to transition state, F is the applied force, kB is the Boltzmann constant, and T is the absolute temperature. The slope on the log plot corresponds to an approximate distance to transition state ∆x, and the intrinsic rate constant k0 can be obtained from the intercept. For example, in 1000 mM KCl (Fig. 4.5c), we obtained ∆xu=1.6 nm and ∆xf =-1.1 nm from the slope of the fit (Table 4.1). ∆xu and ∆xf are the changes in extension to transition state from folded state and unfolded state respectively. Figure 4.6 shows the force dependent rate constants over a range of KCl concentra- tion with Bell model fitting. Using single-molecule optical tweezers, we measured a range of force down to 2 pN. From the salt-dependent data, we observed that the equivalent force is around 2-3 pN in the high salt concentration we measured. The equivalent force F1/2 was defined as the point of intersection of rate constants. At force F1/2, the molecule has equal probability to stay at folded or unfolded states. The equivalent force F1/2 goes up as the increasing salt concentration. It indicates that folded TBA-GQ structure is more mechani- cally stable with more salt. 98 Figure 4.5: Equilibrium measurements of TBA-GQ under force. (a) A representative time trajectory of a single TBA-GQ molecule in constant-position mode at corresponding mean forces (2, 3, 4, and 5 pN) in the presence of 1000 mM KCl. The upper red line and lower yellow line are polymer models for folded and unfolded TBA-GQ respectively. Folded and unfolded dwell times change with forces. (b) Histograms derived from this time trace at different forces are fit with two-state Gaussian distribution. The higher force represents folded state and the lower force represents the unfolded state. The population shifts from folded state to unfolded state as the applied force increases. (c) Force-dependent rate constants of TBA-GQ at 1000 mM KCl. The folding rate constants (cross) and unfolding rate constants (triangle) were fit with Bell model (black dashed lines). 99 Figure 4.6: Force-dependent rate constants of TBA-GQ over a range of KCl concentration with Bell model fitting. Equivalent force F1/2 decreased as KCl concen- tration decreased. The folding rate constants (cross) and unfolding rate constants (triangle). Each KCl concentration was fit with Bell model (black dashed lines). 100 Bell model (force-indep transition extension) kf 0 (1/s) [KCl] 600 mM 0.403 ± 0.075 800 mM 0.549 ± 0.075 1000 mM 0.419 ± 0.078 1500 mM 0.580 ± 0.128 ∆xf (nm) -1.7 ± 0.2 -1.6 ± 0.2 -1.1 ± 0.2 -1.3 ± 0.2 ku0 (1/s) 0.056 ± 0.013 0.098 ± 0.019 0.062 ± 0.014 0.034 ± 0.009 ∆xu (nm) 2.1 ± 0.3 1.5 ± 0.2 1.6 ± 0.3 2.2 ± 0.3 Table 4.1: Bell model fit parameters of rate constants and distances to transition state. ki0 are the rate constants at zero force. ∆xf and ∆xu are the distance to transition state from the unfolded and folded state, respectively. 4.3.2 Salt dependent It well known that monovalent salt ions such as K+ and N a+ are required for stable TBA- GQ structure formation with K+ having the strongest effect [112]. Indeed, we observed that the K+ ion concentration had a strong effect on TBA-GQ mechanical stability. Figure 4.7a shows both the unfolding and refolding rate constants vs K concentration at a mean force of 3 pN (near the force equivalence point for high salt). The unfolding rate constant decreased very slightly with increasing [K+]. However, the folding rate constant showed very strong [K+] dependence and increased strongly with increasing [KCl]. This is consistent with pre- vious investigations and structural models predicting the need for a K+ ion bound between G-quadruplex layers. It indicates that the main effect of K+ ions is helping TBA-GQ for- mation, rather than stabilizing the TBA-GQ structure against mechanical unfolding. The magnitude of the change in both folding and unfolding rate constants at higher salt was similar (approximate 2-fold increase versus decrease for folding vs unfolding rate constant respectively for [K+] increasing from 600 to 1500 mM, at 3 pN force). It was very difficult to observe TBA-GQ folding with only NaCl present even up to 1000 mM NaCl. We thus next added an increasing amount of NaCl to a fixed minimal amount of KCl (300 mM). 101 With increasing [NaCl], we still observed the same two folded and unfolded states. The folding rate constant increased with increasing NaCl, but more modestly compared with KCl (Fig. 4.7b). This suggests that electrostatic screening by monovalent salt only weakly contributes to the increasing folding rate constant with increasing salt, and that the primary K+ concentration dependence indicates the necessity of a bound K+ ion within the TBA- GQ structure. Figure 4.7: Salt effects on TBA-GQ formation. (a) Folding (blue circles) and unfolding (red triangles) rate constants versus KCl concentration at 3 pN. (b) Folding rate constant versus total salt concentration at 3 pN from a range of KCl concentration (blue diamonds) and 300 mM KCl plus a range of NaCl (green crosses). 102 4.3.3 Thrombin dependent We next considered the impact of thrombin on TBA-GQ mechanical stability (Fig. 4.8). These experiments were conducted in a sample chamber containing two side-by-side fluid channels with and without thrombin. Individual tethers were first formed and tested for quality within the blank fluid. The tethers were next transferred to the adjacent thrombin- containing fluid and fixed trap measurements were performed as above. Figure 4.8a shows an example measurement where the tether was in the thrombin solution for 350 seconds and then returned to the adjacent blank channel as a control. While the same two folded and unfolded states were observed, it is clear that the folded state dwell times were much longer with thrombin. Figure 4.8b shows the folding and unfolding rate constants versus thrombin concentration. While thrombin had almost no effect on the folding rate constant, it substantially decreased the unfolding rate constant. Thus, thrombin binding to TBA-GQ does neither induces nor inhibits folding but it does stabilize the already folded TBA-GQ structure. This is opposite the effect of K+. 103 Figure 4.8: Thrombin effects on TBA-GQ resolution. (a) Representative time trajec- tory of thrombin-TBA-GQ complex. Tether was in a buffer containing 150 nM thrombin and then moved to the same buffer without thrombin. (b) Folding (blue circles) and unfolding (red triangles) rate constant versus thrombin concentration at 3 pN. 4.4 Discussion 4.4.1 Bell model Both dwell time analysis and histogram analysis revealed that TBA-GQ is a two-state system. TBA-GQ stays in either folded state or unfolded state with an energy barrier, the transition state, in between. With the applied forces, the free energy surface is tilted by force along the reaction coordinate. For a simple two-state system like TBA-GQ, the basic approach to understand the effect of force on the free energy landscape is Bell model [111] with the 104 assumption that the distance to transition state does not change with forces. Force dependent folding and unfolding rate constants were determined from dwell time as described in Section 4.2. All the dwell times (over hundreds of events for each con- ditions) were used to fit the model by the maximum likelihood global fitting. This method enables us to have enough statistics to determine the fit parameters. The folding and un- folding rate constants at force F depend on the barrier height ∆xf · F and ∆xu · F , where ∆xf and ∆xu are the distances to the transition state along the reaction coordinate from the folded state and unfolded state, respectively. From Fig. 4.6 and Table 4.1, we found a little surprising that the slopes ∆xf and ∆xu are about the same for both folding and unfolding processes. Usually the folding has a larger slope because ∆xf is the distance from an unfolded polymer to a much more compact transition state structure. On the other hand, unfolding is going from a compact folded structure to another compact folded transition state structure, so ∆xu could be much smaller. The similar slope values suggest that the transition state structure is stretchy. 4.4.2 Force-dependent structure diameter model Note that the assumption for Bell model is that the distance is independent of force. It is only true when we consider a small range of forces. In most cases, the force dependent extension must be taken into account especially at low force region due to the large conformation fluctuations. Here, we consider the force-dependent structure diameter model [113]. The distance to transition state is changing with force, therefore the energy barrier ∆x· F in Bell model becomes integral from 0 to F. 105 And the modified force dependent rate constant equations become (cid:90) F 0 ∆x(F ) kBT dF (cid:90) F kf = kf 0 · exp( (cid:90) F ku = ku0 · exp( 0 xt(F ) − xu(F ) kBT dF ) xt(F ) − xN (F ) kBT dF ) (4.2) (4.3) 0 where xt, xu, and xN are the end-to-end extension for the transition state structure, un- folded state structure, and native folded state structure, respectively. The footnote f and u represent folding and unfolding. It is straightforward to estimate the force-dependent extension change of the diameter of the unfolded state structure. The unconstructed 15mer can be described using extensible worm-like chain polymer model. xu(F ) = nssDN A · hssDN A · xXW LC (F ) (4.4) where nssDN A = 15 is the number of free nucleotides, hssDN A = 0.59 nm/nt is the ssDNA unit length, and xXW LC is the extensible worm-like chain extension over contour. For the native folded state structure, we consider it as a rigid body and describe the end-to-end extension using a freely-jointed chain polymer model. xN (F ) = dN · coth( dN · F kBT ) − kBT F (4.5) 106 where dN is the end-to-end extension of the native folded state structure along the direction of force. The dN = 1.5 nm is obtained from the NMR structure (PDB ID: 1C32) [86]. Next, we consider the transition state structure is a structure with the folded portion and unfolded portion. The end-to-end extension is comprised of a rigid body plus a portion of unfolded ssDNA [109]. We again describe the rigid body part using the freely-jointed chain polymer model, and the ssDNA part using the standard extensible worm-like chain polymer model. xt(F ) = dt · coth( dt · F kBT ) − kBT F + nssDN A · hssDN A · xXW LC (F ) (4.6) where the end-to-end extension of the rigid body dt and the number of free nucleotides nssDN A are two fit parameters. We used the same transition state to fit both folding and unfolding force-dependent rate constants because we expect TBA-GQ has a single one-dimensional free energy land- scape. 4.4.3 Finding transition state structure The transition state structure is the key to understand the folding/unfolding free energy landscape. Recent MD simulations from Yang et al. [102] predicted three large barrier- crossing events (i.e. possible transition state structures) during the TBA-GQ formation: the mixed triplex-quadruplex structure (“G-TQ”, Fig. 4.9a), the two-layer G-triplex structure (“G-triplex”, Fig. 4.9b), and the G-hairpin structure (“G-hairpin”, Fig. 4.9c). These three structures contain different number of free nucleotides where nssDN A = 0 for G-TQ, 107 nssDN A = 4 for G-triplex, and nssDN A = 8 for G-hairpin. From our rate constants versus force analysis, we are able to verify these transition state structures by comparing the fit parameter dt to the end-to-end extensions obtained from TBA-GQ crystal structure. Figure 4.9: Schematic possible transition state structures during the folding and unfolding pathway of TBA-GQ. Structures and the corresponding end-to-end diameters of TBA-GQ. (a) G-TQ structure. nssDN A = 0. (b) G-triplex structure. nssDN A = 4. (c) G-hairpin structure. nssDN A = 8. The orange dots represent guanine bases, and the dots with gray rectangles represent the guanines pair by Hoogsteen hydrogen bonds. Considering the polymer model of the transition state structure above, we have one fit parameter, the end-to-end extension dt of the rigid body portion. We used the force- dependent rate constants data in 800 mM KCl buffer as the example, and summarized the fit parameters with different number of free nucleotides (nssDN A = 0, 4, and 8) in transition state in Table 4.2. In the force-dependent structure diameter model, the intrinsic rate constants are similar among all three cases, but the transition state extensions decrease when more free nucleotides are included. When the number of free nucleotides is zero, the whole structure is considered as a rigid body, the transition state extension is 3.3 ± 0.6 nm. It is much larger than the end-to-end extension 1.8 nm in G-TQ structure (Fig. 4.9a). The mixed triplex-quadruplex 108 Force-dep transition extension model # of free nt nssDN A = 0 (G-TQ) nssDN A = 4 (G-triplex) nssDN A = 8 (G-hairpin) kf 0 (1/s) 0.324 ± 0.030 0.341 ± 0.036 0.340 ± 0.019 ku0 (1/s) xt0 (nm) 0.156 ± 0.018 3.3 ± 0.6 0.164 ± 0.015 2.2 ± 0.2 0.163 ± 0.010 0.4 ± 0.1 Table 4.2: Fit parameters of rate constants and transition state diameters with a segment of ssDNA (0, 4, and 8 free nucleotides) in 800 mM KCl buffer. ki0 are the rate constants at zero force. dt are the end-to-end extension of the rigid body portion of the transition state along the applied force. structure does not fit our force-dependent rate constant data. Besides, the transition from the mixed triplex-quadruplex structure to the native folded TBA-GQ is more unlikely because it requires the breaking of the misfolded quadruplex layer and the formation of the correct G-tetrads. The fit extension is also much larger than the size of folded TBA-GQ. We cannot think of any configuration with the diameter 3.3 nm. It indicates that the transition state structure cannot be considered as a rigid body only. When the number of free nucleotides is 4, which corresponds to the G-triplex structure with one 4-nt free polymer at one end, the transition state diameter from the fit is 2.2 ± 0.2 nm. It is close to the diagonal distance 2.1 nm from the G-triplex structure (Fig. 4.9b). Our data is well-represented by the G-triplex model and the transition from the two-layer G-triplex to the native folded state only requires the closing of the free end. When the number of free nucleotides is 8, which corresponds to the G-hairpin structure with two 4-nt free polymers on both ends, the transition state diameter from the fit is 0.4 ± 0.1 nm. It is much smaller than the end-to-end extension 1.5 nm in G-hairpin structure (Fig. 4.9c). These fit results suggest that the G-triplex is the transition state structure for both TBA-GQ folding and unfolding pathways. We used this G-triplex model to fit other salt dependent data. The fit parameters 109 are summarized in Table 4.3. The intrinsic folding rate constant increases and the intrinsic unfolding rate constant decreases with increasing salt concentration. The trend is consistent with Fig. 4.7a, the rate constants versus salt concentration analysis. There is no obvious trend on the transition state distance versus salt concentration. They stay roughly the same diameter which is well-represented by G-triplex model within errors. The red solid lines in Fig. 4.10 show the results of the force-dependent structure diameter model. The rate constant rolls over at low force when force close to 0. Unfolding rate constant at zero force slightly higher than Bell model, and folding rate constant at zero force slightly lower than Bell model. Force-dep transition extension model kf 0 (1/s) [KCl] 600 mM 0.203 ± 0.021 800 mM 0.341 ± 0.036 1000 mM 0.330 ± 0.037 1500 mM 0.354 ± 0.044 ku0 (1/s) 0.130 ± 0.016 0.164 ± 0.015 0.097 ± 0.011 0.077 ± 0.011 xt0 (nm) 2.5 ± 0.2 2.2 ± 0.2 2.5 ± 0.2 2.7 ± 0.2 Table 4.3: Fit parameters of rate constants and transition state position with a range of KCl concentration using the G-triplex plus 4 free nucleotides model. 110 Figure 4.10: Force-dependent rate constants of TBA-GQ over a range of KCl concentration with force-dependent structure diameter model. Equivalent force F1/2 decreased as KCl concentration decreased. The folding rate constants (cross) and unfolding rate constants (triangle) were obtained from 9 (with 303 folding events and 308 unfolding events), 6 (with 645 folding events and 648 unfolding events), 6 (with 377 folding events and 381 unfolding events), and 4 (with 250 folding events and 253 unfolding events) molecules at 600 mM, 800 mM, 1000 mM, and 1500 mM KCl concentration, respectively. Each KCl concentration was fit with Bell model (black dashed lines) and force-dependent structure diameter model (red solid lines). 111 4.4.4 G-triplex as the transition state A proposed TBA-GQ energy landscape model of the one-dimensional reaction coordinate is illustrated in Fig. 4.11. The equilibrium measurement analysis reveals that G-triplex is the major energy barrier for TBA-GQ formation, not G-TQ or G-hairpin. TBA-GQ goes through the same transition state, G-triplex, when it is folded and unfolded. The existence of this two-layer G-triplex structure is supported by other experimental and computational studies [102–104, 114, 115]. Although the G-triplex structure was not stable enough in our study to be characterized as an intermediate state, the truncated form of TBA-GQ, TBA-11 (5’-GGTTGGTGTGG-3’) which the last four nucleotides from TBA-GQ were removed, has been proved to be a stable G-triplex structure by the NMR experiments (PDB ID: 2MKM and 2MKO) [114, 115]. In addition, our thrombin-aptamer experiments also support the G-triplex model. Since two TT loops in the TBA-GQ interact with thrombin exosite I pocket [116], the impact of thrombin on TBA-GQ implies that the TT loops are critical for TBA-GQ unfolding. If two TT loops bind to the hydrophobic pocket in thrombin, it would prevent either terminals peel off from the native folded state. Therefore, the unfolding rate constants decrease. Our results support that the G-quadruplex structure of TBA-GQ is necessary for its interaction with thrombin [89, 116–119], but we did not find that the thrombin promotes TBA-GQ formation as some other groups reported with circular dichroism spectroscopy, fluorescence measurements, and mass spectrometry [120–122]. In those bulk experiments, they measured the fraction of folded TBA-GQ at the present of thrombin. They are not able to distinguish the contribution from increasing folding rate and decreasing unfolding rate. Our results suggest that the fraction folded increased in bulk experiments at the present of thrombin 112 was from the unfolding rate decreased, not the folding rate increased. Since our data show that TBA-GQ is a simple two-state system and the fitting results support the G-triplex structure, we propose the model that the G-triplex is the transition state with one arm peeled from the core and produced a 4-nucleotide ssDNA. TBA-GQ is the simplest G-quadruplex structure. Our study has deepened our understanding on G- quadruplex and paved the way to further investigate the more complex G-quadruplex folding mechanism. Several telomeric G-quadruplex studies also proposed G-triplex as an interme- diate state [123–128]. The complex kinetics indicate the presence of intermediates and the possibility of multi-pathway folding mechanism in telomeric G-quadruplex. Figure 4.11: Proposed reaction diagram explains the folding and unfolding path- way of TBA-GQ. TBA-GQ goes through the same transition state, G-triplex, when it is folded and unfolded. 113 4.5 Conclusion G-quadruplex has attracted attention in recent years because the G-quadruplex structures exist throughout the human genome. The high-throughput sequencing-based method has identified over 700,000 G-quadruplex structures in the human genome [129]. The force- dependent folding and unfolding kinetics studies are essential for understanding the energy landscape. Moreover, the nucleic acid structures like G-quadruplex in cells are under me- chanical force before interacting with proteins (polymerases, helicases, ribosomes, telomerase, etc.) [130]. Single molecule manipulation is a powerful tool to study the folding kinetics of nucleic acids and proteins, and the interaction between nucleic acids and proteins. We can apply the same technique to the G-quadruplex structures in human genome to study their regulatory roles in biological processes. In summary, we have investigated the unfolding and refolding of TBA-GQ at low forces using high-resolution optical tweezers. We have demonstrated that TBA-GQ is a two-state, force-assisted unfolding system. We have identified that the formation of TBA is facilitated by metal ions and the resolution of TBA-GQ is reduced by thrombin. We have also clearly shown that the G-triplex is the transition state between folded and unfolded TBA-GQ structure. These results provide us a better understanding of the energy land- scape of the G-quadruplex and the interaction of the thrombin-aptamer. 114 APPENDICES 115 Appendix A Protocols for performing experiments A.1 DNA construct 3.2 kb dsDNA construct This protocol makes a 3.2 kb dsDNA product for trap performance demonstration in Chap- ter 2. Full construct has a biotin label and a dig label on each end. Materials • Forward primer: /5Biosg/ GAC AGC ATC GCC AGT CAC TA • Reverse primer: /5DigN/ AGC CCT CCC GTA TCG TAG TT • Phusion high-fidelity PCR master mix with HF buffer (NEB M0531S, 2x concentration) • pBR322 DNA template (NEB N3033S) • DMSO (NEB B0515A) • Nuclease-free water (IDT) • IDTE (10 mM Tris and 0.1 mM EDTA) pH 8.0 buffer (IDT) • QIAquick PCR purification kit (Qiagen) 116 Procedure 1. PCR synthesis (a) Prepare one tube for a final volume of 50 µL. (b) Mix 17.5 µL of nuclease-free water, 2.5 µL of 10 µM forward primer, 2.5 µL of 10 µM reverse primer, 1 µL of pBR322 DNA template (10 ng/µL concentration), 1.5 µL of DMSO, and 25 µL of Phusion. (c) Run PCR program for 3.2 kb dsDNA (elongation rate 15 s/kb at 72oC): (1) 98oC for 30 seconds, (2) 98oC for 10 seconds, (3) 59oC for 10 seconds, (4) 72oC for 48 seconds, (5) repeat steps 2-4 for 30 times, (6) 72oC for 5 minutes, (7) hold at 4oC. 2. PCR purification To purify PCR product, follow the QIAquick PCR purification kit. Add 30 uL elution buffer instead of 50 uL to get a higher DNA concentration. 3.2 kb dsDNA construct with short ssDNA insert This protocol makes a ∼3.2 kb dsDNA product with a short single-stranded insert in the middle. The sequence of insert is designed for particular experiments. Three segments: left handle (LH), insert (Insert), and right handle (RH), are synthesized separately and then joined together by ligation. The biotin-labeled LH is a ∼1.5 kb dsDNA with ssDNA over- hang. The digoxigenin-labeled RH is a ∼1.7kb dsDNA with ssDNA overhang. The sequence of insert is extended to anneal to the complementary sequences on both DNA handles. All constructs we used in Chapter 3 and Chapter 4 follow this protocol. For the cryptic binding hairpin construct, slightly different steps are being taken. Please see footnotes. 117 Materials • The sequences of primers, inserts, and probes are listed in Table A.1, A.2, A.3, and A.4. • Phusion high-fidelity PCR master mix with HF buffer (NEB M0531S, 2x concentration) • DNA template pBR322 (NEB N3033S) and lambda (NEB N3011S) • DMSO (NEB B0515A) • Nuclease-free water (IDT) • IDTE (10 mM Tris and 0.1 mM EDTA) pH 8.0 buffer (IDT) • QIAquick PCR purification kit and gel extraction kit (Qiagen) • Restriction enzyme PspGI (NEB R0611S) and TspRI (NEB R0582S) • CutSmart (NEB B7204S) • GelGreen (Biotium 41005) • 1 kb DNA ladder (NEB N3232S) • Gel loading dye (NEB B7025S) • T4 DNA ligase (NEB M0202S) and T4 DNA ligase reaction buffer (B0202S) Procedure 1. PCR synthesis (a) Prepare one tube of a final volume of 50 µL for each handle. 118 (b) For LH: Mix 35 µL of nuclease-free water, 5 µL of 10 µM forward primer, 5 µL of 10 µM reverse primer, 2 µL of pBR322 DNA template (10 ng/µL concentration), 3 µL of DMSO, and 50 µL of Phusion. (c) For RH: Mix 33 µL of nuclease-free water, 5 µL of 10 µM forward primer, 5 µL of 10 µM reverse primer, 4 µL of lambda DNA template (5 ng/µL concentration)¶, 3 µL of DMSO, and 50 µL of Phusion. (d) Run PCR program for 2.2 kb dsDNA (elongation rate 15 s/kb at 72oC) on both reaction mixes: (1) 98oC for 30 seconds, (2) 98oC for 10 seconds, (3) 59oC for 10 seconds, (4) 72oC for 33 seconds, (5) repeat steps 2-4 for 30 times, (6) 72oC for 5 minutes, (7) hold at 4oC. 2. PCR purification (a) To purify PCR products, follow the QIAquick PCR purification kit. Add 30 µL elution buffer instead of 50 µL to get a higher DNA concentration. (b) Measure the PCR products using NanoDrop 2000 spectrophotometer (Thermo Fisher). The typical concentration is 150-200 ng/µL. 3. DNA handle digestion Cut the PCR products to produce the single strand overhangs on handles for ligation. The LH is digested with the restriction enzyme PspGI to produce a 1.5 kb dsDNA with 5-nt overhang on the 5’ end. The RH is digested with the restriction enzyme TspRI to produce a 1.7 kb dsDNA with 9-nt overhang on the 3’ end.(cid:107) ¶use pBR322 DNA template for cryptic binding hairpin construct (cid:107)Do not need RH digestion for cryptic binding hairpin construct 119 (a) Add 3.5 µL of CutSmart buffer and 2 µL of PspGI to the 30 µL LH PCR product. Incubate for 1 hour at 75oC. (b) Add 3.5 µL of CutSmart buffer and 2 µL of TspRI to the 30 µL RH PCR product. Incubate for 1 hour at 65oC. 4. DNA handle purification (a) Make gel. Add 50 mL 0.5x TBE buffer, 0.5 g agarose (low melting point agarose) to a 125 mL Erlenmeyer flask. Microwave until solution just starts to boil and then swirl to mix. Microwave a second time to make sure all gel is dissolved. Cool for 10 minutes, and then add 2 µL GelGreen. Pour gel into a leveled gel box. Remove any bubbles. Place comb into the upper spot. Cover the gel box with foil and wait 2 hours at room temperature to solidify. Put in fridge and use the gel on the same day. (b) Add 3 uL of 100% glycerol and 2 µL loading dye to the digestion products. Preheat RH to 65oC for 1 minute before loading to gel because the 9-nt overhang on RH is too stable at room temperature. (c) Run one lane for each handle. (d) Add a ladder lane (1 µL of 1 kb DNA ladder and 2 µL of gel loading dye to 7 µL nuclease-free water). (e) Run gel for 90 minutes at 75V. (f) Visualize bands with blue light gel imaging system. Cut 1.5 kb LH product and 1.7 kb RH product with a razor blade (Fig. A.1a). Discard the ∼500 bp bands. Typically the digestion yield is 100%. 120 (g) For each handle, put the desired gel slice in a 1.5 mL pre-weighed tube. Get the gel weigh by subtracting tube weight. Usually it is ∼200 mg per gel slice. (h) Follow the QIAquick gel extraction kit for gel extraction. Skip the isopropanol step. To remove the salt from QG buffer, make sure to let the column stand 5 minutes after adding PE buffer. Add 30 µL elution buffer instead of 50 µL to get a higher DNA concentration. Measure the PCR products using NanoDrop. The typical concentration is 50-100 ng/µL. 5. Ligation (a) Calculate the volumes for each components. To achieve a good yield, combine the three components with an equimolar ratio (LH:Insert:RH = 1:1:1). Mass of dsDNA per base pair is 660 g/mol. The mass of a 1.5 kb dsDNA is ∼0.99 g/umol. For a 1.5 kb dsDNA, the concentration 50 ng/µL is close to 50 nM. Typical volumes needed are 25 µL of LH, 25 µL of RH, and 2 µL of Insert. (b) To prepare a final volume of 80 uL reaction mix, add 8 µL of T4 DNA ligase buffer and 4 µL of T4 DNA ligase to the calculated volume of handles and insert, then bring the final volume to 80 µL with nuclease-free water. (c) Run the ligation program: 23oC for 30 minutes then 65oC for 15 minutes. 6. Ligation product purification (a) Make gel (as above). (b) Run gel in two lanes. Put ∼40 µL in each lane. Add a ladder lane. It is also helpful to add a lane of excess unused handle as a reference. 121 (c) The desired final construct band is ∼3 kb (Fig. A.1b). The ∼1.5 kb bands are the leftover handles. Combine gel slices from two lanes in one tube. Follow above steps for gel extraction. Measure the final product with NanoDrop. The typical concentration is 30 ng/µL. (d) Store construct in the fridge. Should be good for 1 month. (a) After digestion. The left lane is the ladder, the middle Figure A.1: Gel purification. lane is the left handle, and the right lane is the right handle. (b) After ligation. The left lane is the ladder, the middle and the right lane is the ligation products. The desired products are the bands in the red boxes. 122 Primer name Sequence (5’ to 3’) /5Biosg/ TGA AGT GGT GGC CTA ACT ACG LH forward primer LH reverse primer TTG CAT GAT AAA GAA GAC AGT CAT RH forward primer RH reverse primer Insert name hp seq Probe name 29mer oligo -bio /5Phos/ TTG AAA TAC CGA CCG CTC AGC TAT CAG CC/idSp/ CTC TGA CAC ATG CAG CTC CC /5DigN/ CAA CAA CGT TGC GCA AAC T Sequence (5’ to 3’) /5Phos/ CCT GG GGC TGA TAG CTG AGC GGT CGG TAT TTC AA A AGT CAA CGT ACT GAT CAC GCT GGA TCC TAG AGT CAA CGT ACT GAT CAC GCT GGA TCC TAT TTT TAG GAT CCA GCG TGA TCA GTA CGT TGA CTC TAG GAT CCA GCG TGA TCA GTA CGT TGA CTT Sequence (5’ to 3’) GG CTG ATA GCT GAG CGG TCG GTA TTT CAA/3Bio/ Table A.1: Complementary hairpin sequence and probe strand for “cryptic bind- ing” experiments. Quantum dots are labeled to the probe strands via biotin-streptavidin bonds. Binding sites are shown in bold. 123 Primer name LH forward primer LH reverse primer RH forward primer RH reverse primer Insert name Sequence (5’ to 3’) /5Biosg/ TGA AGT GGT GGC CTA ACT ACG TTG CAT GAT AAA GAA GAC AGT CAT /5DigN/ GGG CAA ACC AAG ACA GCT AA CCC GTC ATA CAC TTG CTC CT Sequence (5’ to 3’) probe site 1 - 5dT - 5dT /5Phos/CCTGG TTTTT AGG ACT TGT TTTTT CCC probe sites 1 and 2 probe sites 1, 2 and 3 Probe name probe 1 - Cy3 probe 1 - Cy3-Cy5 probe 1 - AF488 probe 2 - Cy5 probe 3 - AF488 ACTGGC /5Phos/CCTGG TTTTT AGG ACT TGT TTTTT CAG TAT CGA TTTTT CCCACTGGC /5Phos/CCTGG TTTTT AGG ACT TGT TTTTT CAG TAT CGA TTTTT ATC GCA AGT TTTTT CCCACT GGC Sequence (5’ to 3’) ACA AGT CCT /3Cy3Sp/ /5Cy5/ ACA AGT CCT /3Cy3Sp/ ACA AGT CCT /3AlexF488N/ TCG ATA CTG /3Cy5Sp/ ACT TGC GAT /3AlexF488N/ Table A.2: ssDNA Insert for synthesizing the “hybridization” construct and the complementary fluorophore-labeled probes. Binding sites are shown in bold. 124 Primer name Sequence (5’ to 3’) /5Biosg/ TGA AGT GGT GGC CTA ACT ACG LH forward primer LH reverse primer TTG CAT GAT AAA GAA GAC AGT CAT RH forward primer RH reverse primer CCC GTC ATA CAC TTG CTC CT /5DigN/ GGG CAA ACC AAG ACA GCT AA Insert name Sequence (5’ to 3’) probe site 1 - 5dT /5Phos/CCTGG TTTTT AGG ACT TGT CCCACTGGC probe site 2 - 5dT /5Phos/CCTGG TTTTT CAG TAT CGA CCCACTGGC 29mer - insert Probe name probe 1 - Cy3 probe 1 - AF488 probe 2 - Cy5 29mer - AF647 /5Phos/CCTGG TTT TT GAA ATA CCG ACC GCT CAG CTA TCA GCC CCCACTGGC Sequence (5’ to 3’) /5Phos/ACA AGT CCT /3Cy3Sp/ /5Phos/ACA AGT CCT /3AlexF488N/ /5Phos/TCG ATA CTG /3Cy5Sp/ /5Phos/GGC TGA TAG CTG AGC GGT CGG TAT TTC AA /3AlexF647N/ Table A.3: ssDNA Insert for synthesizing the “fixed probe” construct and the complementary fluorophore-labeled probes. Binding sites are shown in bold. Primer name Sequence (5’ to 3’) /5Biosg/ TGA AGT GGT GGC CTA ACT ACG LH forward primer LH reverse primer TTG CAT GAT AAA GAA GAC AGT CAT RH forward primer RH reverse primer CCC GTC ATA CAC TTG CTC CT /5DigN/ GGG CAA ACC AAG ACA GCT AA Primer name Sequence (5’ to 3’) /5Biosg/ TGA AGT GGT GGC CTA ACT ACG LH forward primer LH reverse primer TTG CAT GAT AAA GAA GAC AGT CAT RH forward primer RH reverse primer CCC GTC ATA CAC TTG CTC CT /5DigN/ GGG CAA ACC AAG ACA GCT AA Insert name TBA Sequence (5’ to 3’) /5Phos/CCTGG TT GGTTGGTGTGGTTGG TT CCC ACT GGC Table A.4: ssDNA Insert for synthesizing the “TBA-GQ” construct. TBA-GQ sequence is shown in bold 125 A.2 Oxygen scavenger system Materials • Glucose oxidase (Sigma G2133-50KU) • Pyranose oxidase from Coriolus sp. (Sigma P4232-250UN) • Catalase from bovine liver (Sigma C30-100MG) • T50 buffer: 50 mM NaCl and 20 mM Tris-HCl pH8 Procedure 1. In a 0.6 mL centrifuge tube, add 20 µL catalase solution to 80 µL T50 buffer. 2. Add 10 mg glucose oxidase or 5.8 mg pyranose oxidase to the tube. 3. Carefully dissolve by flicking the tube by hand. Do not create foamy bubbles (a possible sign of damage to protein). 4. Centrifuge once at 11 krpm for 5 minutes to remove any possible junk particles (possibly from catalase). Keep the supernatant. 5. Centrifuge filter once at 11 krpm for 1 minute. Keep the flow through. 6. Store the final solution in the fridge. Should last 1-2 weeks. 126 A.3 Anti-blinking system Materials • Trolox (Sigma 238813-1G) • 1M NaOH Procedure 1. Add 5 mL of MilliQ water and 25 µL 1 M NaOH (raises pH to dissolve Trolox) to 15 mL conical tube. 2. Add 5 mg white Trolox powder to the tube to make a 4 mM Trolox solution (near saturated). 3. Wrap tube in foil and rotate (LabQuake) in room temperature for ∼1 hour to dissolve. 4. Vortex and filter everything through a 0.22 um syringe filter. 5. Wrap in foil (light sensitive) and store in the fridge. Should last ∼1 month. 127 Appendix B List of essential parts for the construction of the multi-color fleezers Optical tweezers module • 1064 nm fiber laser, IPG Photonics, YLR-5-1064-LP • AOM1 (acousto-optic modulator), IntraAction, ATM-803D16B • QPD1 & QPD2, First Sensor, QP154-Q-HVSD (1064 nm enhanced quad photodiode) • D1 dichroic mirror, Chroma, T970DCSPXR (short-pass with 1064 nm cutoff) • O1 & O2 (microscope objectives), Nikon, CFI Plan APO VC 60XWI (60x water- immersion, 1.2NA) • Sample chamber stage, Newport, CMA-12CCCL (motor), ESP300 (controller) • D2 dichroic mirror, CVI, SWP-45-RP1064-TU780-PW-1012-C (short-pass with 1064 nm cutoff) • F3 filter, Newport, 10LF25-1064 (1064 nm laser line filter) 128 Three-color confocal fluorescence module • 532 nm laser (green excitation), LASOS, GLK 3250 T01 (LASOS DPSS laser, LasNova series, 50 mW) • 633 nm laser (red excitation), Research Electro-Optics (REO), Red Helium-Neon laser (633 nm, 17 mW) • 473 nm laser (blue excitation), Cobolt, 0473-06-01-0080-100 (MLD series diode laser, 473 nm, 80 mW) • AOM2 & AOM3 (acousto-optic modulator), IntraAction, AOM-802AF1 • D10 dichroic mirror, Semrock, LM01-659-25 (659 nm cutoff, short-pass) • D11 dichroic mirror, Semrock, LM01-552-25 (552 nm cutoff, short-pass) • D12 dichroic mirror, Chroma, ZT473BCM (473 nm cutoff, long-pass) • Single-mode optical fiber, Thorlabs, TS0986175-PM460-HP (PM fiber) • D8 dichroic mirror, Chroma, ZT473BCM (for blue laser) • D9 dichroic mirror, Semrock, LM01-552-25 (for green laser) • PD1, PD2, and PD3 (Si photodetector), Throlabs, PDA36A (PD1 for blue, PD2 for green, and PD3 for red) • D7 multi dichroic mirror, Chroma, ZT473/532/633RPC-XT • Piezo mirror stage, Mad City Labs, Nano-MTA2 Invar • D3 dichroic mirror, Semrock, FF875-Di01-25-D 129 • D4 dichroic mirror, Thorlabs, DMLP900 (long-pass, 900 nm cutoff) • PSD (position-sensitive detector), First Sensor, DL100-7-PCBA3 • F4 filter, Chroma, ZET473/532/635M • PH (pinhole), Thorlabs or Edmund, Precision pinholes 20-100 um • D5 dichroic mirror, Chroma, 540DCXR • F5 filter, Chroma, ET510/20M • D6 dichroic mirror, Chroma, 640DCXR • F6 filter, Chroma, HQ580/60M/2P • F7 filter, Chroma, ET685/50M • APDs (avalanche photodiodes), Excelitas Tech, SPCM-AQRH-14 Bright-field imaging system • LED, Thorlabs, LED940E • F1 filter, Thorlabs, FES1000 (short-pass filter, 1000 nm cutoff) • F2 filter, Thorlabs, FEL850 (long-pass filter, 850 nm cutoff) • CCD camera, Watec, WAT-902H2 130 BIBLIOGRAPHY 131 BIBLIOGRAPHY [1] Francis Crick. Central dogma of molecular biology. Nature, 227(5258):561–563, 1970. [2] Encode Project Consortium. An integrated encyclopedia of dna elements in the human genome. Nature, 489(7414):57–74, 2012. [3] J. D. Watson and F. H. C. Crick. Molecular structure of nucleic acids: A structure for deoxyribose nucleic acid. Nature, 171(4356):737–738, 1953. [4] A. Rich and J. D. Watson. Some relations between dna and rna. Proceedings of the National Academy of Sciences, 40(8):759–764, 1954. [5] S. B. Smith, L. Finzi, and C. Bustamante. Direct mechanical measurements of the elasticity of single dna molecules by using magnetic beads. Science, 258(5085):1122– 1126, 1992. [6] C. Bouchiat, M. D. Wang, J. Allemand, T. Strick, S. M. Block, and V. Croquette. Estimating the persistence length of a worm-like chain molecule from force-extension measurements. Biophys J, 76(1 Pt 1):409–13, 1999. [7] C. Bustamante, J. F. Marko, E. D. Siggia, and S. Smith. Entropic elasticity of lambda- phage dna. Science, 265(5178):1599–600, 1994. [8] S. B. Smith, Y. Cui, and C. Bustamante. Overstretching b-dna: The elastic response of individual double-strandedand single-stranded dna molecules. Science, 271(5250):795– 799, 1996. [9] Lisa J. Lapidus, Peter J. Steinbach, William A. Eaton, Attila Szabo, and James Hofrichter. Effects of chain stiffness on the dynamics of loop formation in polypeptides. appendix: Testing a 1-dimensional diffusion model for peptide dynamics. The Journal of Physical Chemistry B, 106(44):11628–11640, 2002. [10] P. Cluzel, A. Lebrun, C. Heller, R. Lavery, J. L. Viovy, D. Chatenay, and F. Caron. Dna: An extensible molecule. Science, 271(5250):792–4, 1996. [11] M. D. Wang, H. Yin, R. Landick, J. Gelles, and S. M. Block. Stretching dna with optical tweezers. Biophysical Journal, 72(3):1335–1346, 1997. [12] Eric Westhof and Valrie Fritsch. Rna folding: beyond watsoncrick pairs. Structure, 8(3):R55–R65, 2000. 132 [13] K. Hoogsteen. The crystal and molecular structure of a hydrogen-bonded complex between 1-methylthymine and 9-methyladenine. Acta Crystallographica, 16(9):907– 916, 1963. [14] Richard H. Shafer and Ivan Smirnov. Biological aspects of dna/rna quadruplexes. Biopolymers, 56(3):209–227, 2000. [15] D. Rhodes and H. J. Lipps. G-quadruplexes and their regulatory roles in biology. Nucleic Acids Res, 43(18):8627–37, 2015. [16] A. Risitano and K. R. Fox. Influence of loop size on the stability of intramolecular dna quadruplexes. Nucleic Acids Res, 32(8):2598–606, 2004. [17] P. Hazel, J. Huppert, S. Balasubramanian, and S. Neidle. Loop-length-dependent folding of g-quadruplexes. J Am Chem Soc, 126(50):16405–15, 2004. [18] N. Smargiasso, F. Rosu, W. Hsia, P. Colson, E. S. Baker, M. T. Bowers, E. De Pauw, and V. Gabelica. G-quadruplex dna assemblies: Loop length, cation identity, and multimer formation. J Am Chem Soc, 130(31):10208–16, 2008. [19] J. L. Huppert. Structure, location and interactions of g-quadruplexes. FEBS J, 277(17):3452–8, 2010. [20] A. Guedin, J. Gros, P. Alberti, and J. L. Mergny. How long is too long? effects of loop size on g-quadruplex stability. Nucleic Acids Res, 38(21):7858–68, 2010. [21] B. Heddi and A. T. Phan. Structure of human telomeric dna in crowded solution. J Am Chem Soc, 133(25):9824–33, 2011. [22] A. D. Ellington and J. W. Szostak. In vitro selection of rna molecules that bind specific ligands. Nature, 346(6287):818–22, 1990. [23] A. D. Ellington and J. W. Szostak. Selection in vitro of single-stranded dna molecules that fold into specific ligand-binding structures. Nature, 355(6363):850–2, 1992. [24] C. Tuerk and L. Gold. Systematic evolution of ligands by exponential enrichment: Rna ligands to bacteriophage t4 dna polymerase. Science, 249(4968):505–10, 1990. [25] C. Platella, C. Riccardi, D. Montesarchio, G. N. Roviello, and D. Musumeci. G- quadruplex-based aptamers against protein targets in therapy and diagnostics. Biochim Biophys Acta Gen Subj, 1861(5 Pt B):1429–1447, 2017. [26] L. C. Bock, L. C. Griffin, J. A. Latham, E. H. Vermaas, and J. J. Toole. Selec- tion of single-stranded dna molecules that bind and inhibit human thrombin. Nature, 355(6360):564–6, 1992. 133 [27] S. M. Nimjee, C. P. Rusconi, and B. A. Sullenger. Aptamers: an emerging class of therapeutics. Annu Rev Med, 56:555–83, 2005. [28] P.R. Selvin and T. Ha. Single-molecule Techniques: A Laboratory Manual. Cold Spring Harbor Laboratory Press, 2008. [29] P. Hinterdorfer and Y. F. Dufrene. Detection and localization of single molecular recognition events using atomic force microscopy. Nat Methods, 3(5):347–55, 2006. [30] J. R. Moffitt, Y. R. Chemla, S. B. Smith, and C. Bustamante. Recent advances in optical tweezers. Annu Rev Biochem, 77:205–28, 2008. [31] I. De Vlaminck and C. Dekker. Recent advances in magnetic tweezers. Annu Rev Biophys, 41:453–72, 2012. [32] E. Haustein and P. Schwille. Fluorescence correlation spectroscopy: novel variations of an established technique. Annu Rev Biophys Biomol Struct, 36:151–69, 2007. [33] T. Funatsu, Y. Harada, M. Tokunaga, K. Saito, and T. Yanagida. Imaging of sin- gle fluorescent molecules and individual atp turnovers by single myosin molecules in aqueous solution. Nature, 374(6522):555–9, 1995. [34] R. M. Power and J. Huisken. A guide to light-sheet fluorescence microscopy for mul- tiscale imaging. Nat Methods, 14(4):360–373, 2017. [35] Thomas A. Klar and Stefan W. Hell. Subdiffraction resolution in far-field fluorescence microscopy. Optics Letters, 24(14), 1999. [36] M. Higa, M. Fujita, and K. Yoshida. Dna replication origins and fork progression at mammalian telomeres. Genes (Basel), 8(4), 2017. [37] A. Ashkin. Acceleration and trapping of particles by radiation pressure. Physical Review Letters, 24(4):156, 1970. [38] A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and S. Chu. Observation of a single-beam gradient force optical trap for dielectric particles. Opt Lett, 11(5):288, 1986. [39] J. R. Moffitt, Y. R. Chemla, D. Izhaky, and C. Bustamante. Differential detection of dual traps improves the spatial resolution of optical tweezers. Proc Natl Acad Sci U S A, 103(24):9006–11, 2006. [40] K. C. Neuman and S. M. Block. Optical trapping. Rev Sci Instrum, 75(9):2787–809, 2004. 134 [41] M. T. Valentine, N. R. Guydosh, B. Gutierrez-Medina, A. N. Fehr, J. O. Andreasson, and S. M. Block. Precision steering of an optical trap by electro-optic deflection. Opt Lett, 33(6):599–601, 2008. [42] M. S. Woody, M. Capitanio, E. M. Ostap, and Y. E. Goldman. Electro-optic deflectors deliver advantages over acousto-optical deflectors in a high resolution, ultra-fast force- clamp optical trap. Opt Express, 26(9):11181–11193, 2018. [43] C. Gosse and V. Croquette. Magnetic tweezers: Micromanipulation and force mea- surement at the molecular level. Biophysical Journal, 82(6):3314–3329, 2002. [44] K. Visscher, G. J. Brakenhoff, and J. J. Krol. Micromanipulation by ”multiple” optical traps created by a single fast scanning trap integrated with the bilateral confocal scanning laser microscope. Cytometry, 14(2):105–14, 1993. [45] F. Gittes and C. F. Schmidt. Interference model for back-focal-plane displacement detection in optical tweezers. Opt Lett, 23(1):7–9, 1998. [46] R. Roy, S. Hohng, and T. Ha. A practical guide to single-molecule fret. Nat Methods, 5(6):507–16, 2008. [47] M. A. Dijk, L. C. Kapitein, Jv Mameren, C. F. Schmidt, and E. J. Peterman. Com- bining optical trapping and single-molecule fluorescence spectroscopy: enhanced pho- tobleaching of fluorophores. J Phys Chem B, 108(20):6479–84, 2004. [48] S. Hohng, R. Zhou, M. K. Nahas, J. Yu, K. Schulten, D. M. Lilley, and T. Ha. Fluorescence-force spectroscopy maps two-dimensional reaction landscape of the holl- iday junction. Science, 318(5848):279–83, 2007. [49] R. R. Brau, P. B. Tarsa, J. M. Ferrer, P. Lee, and M. J. Lang. Interlaced optical force- fluorescence measurements for single molecule biophysics. Biophys J, 91(3):1069–77, 2006. [50] M. J. Comstock, T. Ha, and Y. R. Chemla. Ultrahigh-resolution optical trap with single-fluorophore sensitivity. Nat Methods, 8(4):335–40, 2011. [51] T. T. Ngo, Q. Zhang, R. Zhou, J. G. Yodh, and T. Ha. Asymmetric unwrapping of nucleosomes under tension directed by dna local flexibility. Cell, 160(6):1135–44, 2015. [52] M. J. Comstock, K. D. Whitley, H. Jia, J. Sokoloski, T. M. Lohman, T. Ha, and Y. R. Chemla. Direct observation of structure-function relationship in a nucleic acid- processing enzyme. Science, 348(6232):352–4, 2015. 135 [53] S. Suksombat, R. Khafizov, A. G. Kozlov, T. M. Lohman, and Y. R. Chemla. Struc- tural dynamics of e. coli single-stranded dna binding protein reveal dna wrapping and unwrapping pathways. Elife, 4, 2015. [54] V. K. Duesterberg, I. T. Fischer-Hwang, C. F. Perez, D. W. Hogan, and S. M. Block. Observation of long-range tertiary interactions during ligand binding by the tpp ri- boswitch aptamer. Elife, 4, 2015. [55] A. G. Baker, C. Y. Chuang, M. Whitmore, and M. J. Comstock. Randomizing phase to remove acousto-optic device wiggle errors for high-resolution optical tweezers. Appl Opt, 57(8):1752–1756, 2018. [56] E. A. Abbondanzieri, W. J. Greenleaf, J. W. Shaevitz, R. Landick, and S. M. Block. Direct observation of base-pair stepping by rna polymerase. Nature, 438(7067):460–5, 2005. [57] T. T. Perkins. Angstrom-precision optical traps and applications. Annu Rev Biophys, 43:279–302, 2014. [58] K. Berg-Sorensen and H. Flyvbjerg. Power spectrum analysis for optical tweezers. Review of Scientific Instruments, 75(3):594–612, 2004. [59] P. Tinnefeld and M. Sauer. Branching out of single-molecule fluorescence spec- troscopy: challenges for chemistry and influence on biology. Angew Chem Int Ed Engl, 44(18):2642–71, 2005. [60] M. J. Lang, P. M. Fordyce, A. M. Engh, K. C. Neuman, and S. M. Block. Simultaneous, coincident optical trapping and single-molecule fluorescence. Nat Methods, 1(2):133–9, 2004. [61] Y. He, M. Lu, J. Cao, and H. P. Lu. Manipulating protein conformations by single- molecule afm-fret nanoscopy. ACS Nano, 6(2):1221–9, 2012. [62] G. Sirinakis, Y. Ren, Y. Gao, Z. Xi, and Y. Zhang. Combined versatile high- resolution optical tweezers and single-molecule fluorescence microscopy. Rev Sci In- strum, 83(9):093708, 2012. [63] A. L. Forget, C. C. Dombrowski, I. Amitani, and S. C. Kowalczykowski. Exploring protein-dna interactions in 3d using in situ construction, manipulation and visual- ization of individual dna dumbbells with optical traps, microfluidics and fluorescence microscopy. Nat Protoc, 8(3):525–38, 2013. [64] I. Heller, G. Sitters, O. D. Broekmans, G. Farge, C. Menges, W. Wende, S. W. Hell, E. J. Peterman, and G. J. Wuite. Sted nanoscopy combined with optical tweezers reveals protein dynamics on densely covered dna. Nat Methods, 10(9):910–6, 2013. 136 [65] S. Lee and S. Hohng. An optical trap combined with three-color fret. J Am Chem Soc, 135(49):18260–3, 2013. [66] F. E. Kemmerich, M. Swoboda, D. J. Kauert, M. S. Grieb, S. Hahn, F. W. Schwarz, R. Seidel, and M. Schlierf. Simultaneous single-molecule force and fluorescence sampling of dna nanostructure conformations using magnetic tweezers. Nano Lett, 16(1):381–6, 2016. [67] M. P. Landry, P. M. McCall, Z. Qi, and Y. R. Chemla. Characterization of photoacti- vated singlet oxygen damage in single-molecule optical trap experiments. Biophys J, 97(8):2128–36, 2009. [68] M. Swoboda, J. Henig, H. M. Cheng, D. Brugger, D. Haltrich, N. Plumere, and M. Schlierf. Enzymatic oxygen scavenging for photostability without ph drop in single- molecule experiments. ACS Nano, 6(7):6364–9, 2012. [69] T. Ha and P. Tinnefeld. Photophysics of fluorescent probes for single-molecule bio- physics and super-resolution imaging. Annu Rev Phys Chem, 63:595–617, 2012. [70] L. R. Brewer and P. R. Bianco. Laminar flow cells for single-molecule studies of dna- protein interactions. Nat Methods, 5(6):517–25, 2008. [71] C. G. Baumann, S. B. Smith, V. A. Bloomfield, and C. Bustamante. Ionic effects on the elasticity of single dnamolecules. Proc Natl Acad Sci U S A, 94(12):6185–90, 1997. [72] J. van Mameren, G. J. Wuite, and I. Heller. Introduction to optical tweezers: back- ground, system designs, and commercial solutions. Methods Mol Biol, 783:1–20, 2011. [73] A. P. Alivisatos, W. Gu, and C. Larabell. Quantum dots as cellular probes. Annu Rev Biomed Eng, 7:55–76, 2005. [74] G. Mustafa, C. Y. Chuang, W. A. Roy, M. M. Farhath, N. Pokhrel, Y. Ma, K. Naga- sawa, E. Antony, M. J. Comstock, S. Basu, and H. Balci. A force sensor that converts fluorescence signal into force measurement utilizing short looped dna. Biosens Bio- electron, 121:34–40, 2018. [75] M. Gellert, M. N. Lipsett, and D. R. Davies. Helix formation by guanylic acid. Proceed- ings of the National Academy of Sciences of the United States of America, 48(12):2013– 2018, 1962. [76] C. K. Kwok and C. J. Merrick. G-quadruplexes: Prediction, characterization, and biological application. Trends Biotechnol, 35(10):997–1013, 2017. 137 [77] R. Hansel-Hertsch, M. Di Antonio, and S. Balasubramanian. Dna g-quadruplexes in the human genome: detection, functions and therapeutic potential. Nat Rev Mol Cell Biol, 18(5):279–284, 2017. [78] E. Y. Lam, D. Beraldi, D. Tannahill, and S. Balasubramanian. G-quadruplex structures are stable and detectable in human genomic dna. Nat Commun, 4:1796, 2013. [79] R. Rodriguez, K. M. Miller, J. V. Forment, C. R. Bradshaw, M. Nikan, S. Britton, T. Oelschlaegel, B. Xhemalce, S. Balasubramanian, and S. P. Jackson. Small-molecule- induced dna damage identifies alternative dna structures in human genes. Nat Chem Biol, 8(3):301–10, 2012. [80] A. M. Zahler, J. R. Williamson, T. R. Cech, and D. M. Prescott. Inhibition of telom- erase by g-quartet dna structures. Nature, 350(6320):718–20, 1991. [81] S. Balasubramanian, L. H. Hurley, and S. Neidle. Targeting g-quadruplexes in gene promoters: a novel anticancer strategy? Nat Rev Drug Discov, 10(4):261–75, 2011. [82] S. A. Ohnmacht, C. Marchetti, M. Gunaratnam, R. J. Besser, S. M. Haider, G. Di Vita, H. L. Lowe, M. Mellinas-Gomez, S. Diocou, M. Robson, J. Sponer, B. Islam, R. B. Pedley, J. A. Hartley, and S. Neidle. A g-quadruplex-binding compound showing anti- tumour activity in an in vivo model for pancreatic cancer. Sci Rep, 5:11385, 2015. [83] S. Burge, G. N. Parkinson, P. Hazel, A. K. Todd, and S. Neidle. Quadruplex dna: sequence, topology and structure. Nucleic Acids Res, 34(19):5402–15, 2006. [84] H. Han and L. H. Hurley. G-quadruplex dna: a potential target for anti-cancer drug design. Trends Pharmacol Sci, 21(4):136–42, 2000. [85] R. F. Macaya, P. Schultze, F. W. Smith, J. A. Roe, and J. Feigon. Thrombin-binding dna aptamer forms a unimolecular quadruplex structure in solution. Proceedings of the National Academy of Sciences, 90(8):3745–3749, 1993. [86] V. M. Marathias and P. H. Bolton. Structures of the potassium-saturated, 2:1, and intermediate, 1:1, forms of a quadruplex dna. Nucleic Acids Research, 28(9):1969–1977, 2000. [87] D. Musumeci and D. Montesarchio. Polyvalent nucleic acid aptamers and modulation of their activity: a focus on the thrombin binding aptamer. Pharmacol Ther, 136(2):202– 15, 2012. [88] L. C. Griffin, G. F. Tidmarsh, L. C. Bock, J. J. Toole, and L. L. Leung. In vivo antico- agulant properties of a novel nucleotide-based thrombin inhibitor and demonstration of regional anticoagulation in extracorporeal circuits. Blood, 81(12):3271–6, 1993. 138 [89] K. Padmanabhan, K. P. Padmanabhan, J. D. Ferrara, J. E. Sadler, and A. Tulinsky. The structure of alpha-thrombin inhibited by a 15-mer single-stranded dna aptamer. J Biol Chem, 268(24):17651–4, 1993. [90] K. Y. Wang, S. Mccurdy, R. G. Shea, S. Swaminathan, and P. H. Bolton. A dna aptamer which binds to and inhibits thrombin exhibits a new structural motif for dna. Biochemistry, 32(8):1899–1904, 1993. [91] M. Vairamani and M. L. Gross. G-quadruplex formation of thrombin-binding aptamer detected by electrospray ionization mass spectrometry. J Am Chem Soc, 125(1):42–3, 2003. [92] B. Pagano, L. Martino, A. Randazzo, and C. Giancola. Stability and binding properties of a modified thrombin binding aptamer. Biophys J, 94(2):562–9, 2008. [93] C. M. Olsen, H. T. Lee, and L. A. Marky. Unfolding thermodynamics of intramolec- ular g-quadruplexes: Base sequence contributions of the loops. J Phys Chem B, 113(9):2587–95, 2009. [94] A. Avino, C. Fabrega, M. Tintore, and R. Eritja. Thrombin binding aptamer, more than a simple aptamer: Chemically modified derivatives and biomedical applications. Current Pharmaceutical Design, 18(14):2036–2047, 2012. [95] S. De Tito, F. Morvan, A. Meyer, J. J. Vasseur, A. Cummaro, L. Petraccone, B. Pagano, E. Novellino, A. Randazzo, C. Giancola, and D. Montesarchio. Fluores- cence enhancement upon g-quadruplex folding: synthesis, structure, and biophysical characterization of a dansyl/cyclodextrin-tagged thrombin binding aptamer. Bioconjug Chem, 24(11):1917–27, 2013. [96] B. Deng, Y. Lin, C. Wang, F. Li, Z. Wang, H. Zhang, X. F. Li, and X. C. Le. Aptamer binding assays for proteins: the thrombin example–a review. Anal Chim Acta, 837:1– 15, 2014. [97] O. Tatarinova, V. Tsvetkov, D. Basmanov, N. Barinov, I. Smirnov, E. Timofeev, D. Kaluzhny, A. Chuvilin, D. Klinov, A. Varizhuk, and G. Pozmogova. Comparison of the ’chemical’ and ’structural’ approaches to the optimization of the thrombin-binding aptamer. PLoS One, 9(2):e89383, 2014. [98] L. J. Aaldering, V. Poongavanam, N. Langkjaer, N. A. Murugan, P. T. Jorgensen, J. Wengel, and R. N. Veedu. Development of an efficient g-quadruplex-stabilised thrombin-binding aptamer containing a three-carbon spacer molecule. Chembiochem, 18(8):755–763, 2017. [99] I. Russo Krauss, V. Napolitano, L. Petraccone, R. Troisi, V. Spiridonova, C. A. Mattia, and F. Sica. Duplex/quadruplex oligonucleotides: Role of the duplex domain in the 139 stabilization of a new generation of highly effective anti-thrombin aptamers. Int J Biol Macromol, 107(Pt B):1697–1705, 2018. [100] B. Basnar, R. Elnathan, and I. Willner. Following aptamer-thrombin binding by force measurements. Anal Chem, 78(11):3638–42, 2006. [101] J. W. Shim, Q. Tan, and L. Q. Gu. Single-molecule detection of folding and unfolding of the g-quadruplex aptamer in a nanopore nanocavity. Nucleic Acids Res, 37(3):972–82, 2009. [102] C. Yang, M. Kulkarni, M. Lim, and Y. Pak. Insilico direct folding of thrombin-binding aptamer g-quadruplex at all-atom level. Nucleic Acids Res, 45(22):12648–12656, 2017. [103] Y. Bian, F. Song, Z. Cao, L. Zhao, J. Yu, X. Guo, and J. Wang. Fast-folding pathways of the thrombin-binding aptamer g-quadruplex revealed by a markov state model. Biophys J, 114(7):1529–1538, 2018. [104] X. Zeng, L. Zhang, X. Xiao, Y. Jiang, Y. Guo, X. Yu, X. Pu, and M. Li. Unfolding mechanism of thrombin-binding aptamer revealed by molecular dynamics simulation and markov state model. Sci Rep, 6:24065, 2016. [105] K. D. Whitley, M. J. Comstock, and Y. R. Chemla. High-resolution optical tweezers combined with single-molecule confocal microscopy. Methods Enzymol, 582:137–169, 2017. [106] John F. Marko and Eric D. Siggia. Stretching dna. Macromolecules, 28(26):8759–8770, 1995. [107] T. Odijk. Stiff chains and filaments under tension. Macromolecules, 28(20):7016–7018, 1995. [108] D. Izadi, Y. Chen, M. L. Whitmore, J. D. Slivka, K. Ching, L. J. Lapidus, and M. J. Comstock. Combined force ramp and equilibrium high-resolution investigations reveal multipath heterogeneous unfolding of protein g. J Phys Chem B, 122(49):11155–11165, 2018. [109] G. Yuan, S. Le, M. Yao, H. Qian, X. Zhou, J. Yan, and H. Chen. Elasticity of the tran- sition state leading to an unexpected mechanical stabilization of titin immunoglobulin domains. Angew Chem Int Ed Engl, 56(20):5490–5493, 2017. [110] B. Jagannathan, P. J. Elms, C. Bustamante, and S. Marqusee. Direct observation of a force-induced switch in the anisotropic mechanical unfolding pathway of a protein. Proc Natl Acad Sci U S A, 109(44):17820–5, 2012. 140 [111] G. Bell. Models for the specific adhesion of cells to cells. Science, 200(4342):618–627, 1978. [112] E. S. Hong, H. J. Yoon, B. Kim, Y. H. Yim, H. Y. So, and S. K. Shin. Mass spec- trometric studies of alkali metal ion binding on thrombin-binding aptamer dna. J Am Soc Mass Spectrom, 21(7):1245–55, 2010. [113] O. K. Dudko, G. Hummer, and A. Szabo. Theory, analysis, and interpretation of single- molecule force spectroscopy experiments. Proc Natl Acad Sci U S A, 105(41):15755–60, 2008. [114] L. Cerofolini, J. Amato, A. Giachetti, V. Limongelli, E. Novellino, M. Parrinello, M. Fragai, A. Randazzo, and C. Luchinat. G-triplex structure and formation propen- sity. Nucleic Acids Res, 42(21):13393–404, 2014. [115] V. Limongelli, S. De Tito, L. Cerofolini, M. Fragai, B. Pagano, R. Trotta, S. Cosconati, L. Marinelli, E. Novellino, I. Bertini, A. Randazzo, C. Luchinat, and M. Parrinello. The g-triplex dna. Angew Chem Int Ed Engl, 52(8):2269–73, 2013. [116] I. Russo Krauss, A. Merlino, A. Randazzo, E. Novellino, L. Mazzarella, and F. Sica. High-resolution structures of two complexes between thrombin and thrombin-binding aptamer shed light on the role of cations in the aptamer inhibitory activity. Nucleic Acids Res, 40(16):8119–28, 2012. [117] J. A. Kelly, J. Feigon, and T. O. Yeates. Reconciliation of the x-ray and nmr structures of the thrombin-binding aptamer d(ggttggtgtggttgg). J Mol Biol, 256(3):417–22, 1996. [118] K. Padmanabhan and A. Tulinsky. An ambiguous structure of a dna 15-mer thrombin complex. Acta Crystallogr D Biol Crystallogr, 52(Pt 2):272–82, 1996. [119] R. Dolot, C. H. Lam, M. Sierant, Q. Zhao, F. W. Liu, B. Nawrot, M. Egli, and X. Yang. Crystal structures of thrombin in complex with chemically modified thrombin dna aptamers reveal the origins of enhanced affinity. Nucleic Acids Res, 46(9):4819–4830, 2018. [120] S. Nagatoishi, Y. Tanaka, and K. Tsumoto. Circular dichroism spectra demonstrate formation of the thrombin-binding dna aptamer g-quadruplex under stabilizing-cation- deficient conditions. Biochem Biophys Res Commun, 352(3):812–7, 2007. [121] E. Baldrich and C. K. O’Sullivan. Ability of thrombin to act as molecular chaper- one, inducing formation of quadruplex structure of thrombin-binding aptamer. Anal Biochem, 341(1):194–7, 2005. 141 [122] J. Zhang, R. R. O. Loo, and J. A. Loo. Structural characterization of a thrombin- aptamer complex by high resolution native top-down mass spectrometry. J Am Soc Mass Spectrom, 28(9):1815–1822, 2017. [123] T. Mashimo, H. Yagi, Y. Sannohe, A. Rajendran, and H. Sugiyama. Folding pathways of human telomeric type-1 and type-2 g-quadruplex structures. J Am Chem Soc, 132(42):14910–8, 2010. [124] D. Koirala, T. Mashimo, Y. Sannohe, Z. Yu, H. Mao, and H. Sugiyama. Intramolecular folding in three tandem guanine repeats of human telomeric dna. Chem Commun (Camb), 48(14):2006–8, 2012. [125] W. Li, X. M. Hou, P. Y. Wang, X. G. Xi, and M. Li. Direct measurement of sequential folding pathway and energy landscape of human telomeric g-quadruplex structures. J Am Chem Soc, 135(17):6423–6, 2013. [126] P. Stadlbauer, M. Krepl, 3rd Cheatham, T. E., J. Koca, and J. Sponer. Structural dynamics of possible late-stage intermediates in folding of quadruplex dna studied by molecular simulations. Nucleic Acids Res, 41(14):7128–43, 2013. [127] R. D. Gray, J. O. Trent, and J. B. Chaires. Folding and unfolding pathways of the human telomeric g-quadruplex. J Mol Biol, 426(8):1629–50, 2014. [128] X. M. Hou, Y. B. Fu, W. Q. Wu, L. Wang, F. Y. Teng, P. Xie, P. Y. Wang, and X. G. Involvement of g-triplex and g-hairpin in the multi-pathway folding of human Xi. telomeric g-quadruplex. Nucleic Acids Res, 45(19):11401–11412, 2017. [129] V. S. Chambers, G. Marsico, J. M. Boutell, M. Di Antonio, G. P. Smith, and S. Bal- asubramanian. High-throughput sequencing of dna g-quadruplex structures in the human genome. Nat Biotechnol, 33(8):877–81, 2015. [130] C. Bustamante, Y. R. Chemla, N. R. Forde, and D. Izhaky. Mechanical processes in biochemistry. Annu Rev Biochem, 73:705–48, 2004. 142