Edited by Igor Medintz Niko Hildebrandt FRET - Förster Resonance Energy Transfer (also available in digital formats) Balzani, V., Ceroni, P., Juris, A. Sauer, M., Hofkens, J., Enderlein, J. Photochemistry and Photophysics Handbook of Fluorescence Spectroscopy and Imaging Concepts, Research Topics, Applications From Single Molecules to Ensembles 2014 ISBN: 978-3-527-33479-7 2011 ISBN: 978-3-527-31669-4 Kubitscheck, U. (ed.) Bräuchle, C., Lamb, D. C., Michaelis, J. (eds.) Fluorescence Microscopy 2013 ISBN: 978-3-527-32922-9 Single Particle Tracking and Single Molecule Energy Transfer Valeur, B., Berberan-Santos, M. N. 2010 ISBN: 978-3-527-32296-1 From Principles to Biological Applications Molecular Fluorescence Principles and Applications Second edition 2012 ISBN (Hardcover): 978-3-527-32837-6 ISBN (Softcover): 978-3-527-32846-8 Yanagida, T., Ishii, Y. (eds.) Single Molecule Dynamics in Life Science 2008 ISBN: 978-3-527-31288-7 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Related Titles FRET - Förster Resonance Energy Transfer From Theory to Applications Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Edited by Igor Medintz and Niko Hildebrandt Dr. Igor Medintz U.S. Naval Research Laboratory Center for Bio/Molecular Science and Engineering 4555 Overlook Avenue, SW Washington D.C. 20375 United States Dr. Niko Hildebrandt Universite Paris-Sud Institut d’Electronique Fondamentale NanoBioPhotonics Bâtiment 220 91405 Orsay Cedex France All books published by Wiley-VCH are carefully produced. Nevertheless, authors, editors, and publisher do not warrant the information contained in these books, including this book, to be free of errors. Readers are advised to keep in mind that statements, data, illustrations, procedural details or other items may inadvertently be inaccurate. Library of Congress Card No.: applied for British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library. Bibliographic information published by the Deutsche Nationalbibliothek The Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie; detailed bibliographic data are available on the Internet at <http://dnb.d-nb.de>. # 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Boschstr. 12, 69469 Weinheim, Germany All rights reserved (including those of translation into other languages). No part of this book may be reproduced in any form – by photoprinting, microfilm, or any other means – nor transmitted or translated into a machine language without written permission from the publishers. Registered names, trademarks, etc. used in this book, even when not specifically marked as such, are not to be considered unprotected by law. Print ISBN: ePDF ISBN: ePub ISBN: Mobi ISBN: oBook ISBN: Cover Design Typesetting 978-3-527-32816-1 978-3-527-65605-9 978-3-527-65604-2 978-3-527-65603-5 978-3-527-65602-8 Adam-Design, Weinheim, Germany Thomson Digital, Noida, India Printing and Binding Singapore Markono Print Media Pte Ltd, Printed on acid-free paper Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Editors Contents Preface XV List of Contributors XIX Part One Background, Theory, and Concepts 1 1 How I Remember Theodor F€ orster 3 Herbert Dreeskamp 2 Remembering Robert Clegg and Elizabeth Jares-Erijman and Their Contributions to FRET 9 Thomas M. Jovin Biographical Sketch of Bob Clegg 10 Biographical Sketch of Eli Jares-Erijman 11 The Pervasive Influence of Gregorio Weber 12 Contributions by Bob Clegg to FRET 12 Contributions by Eli Jares-Erijman to FRET 16 A Final Thought 18 References 19 2.1 2.2 2.3 2.4 2.5 2.6 3 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 3.10 3.11 F€ orster Theory 23 B. Wieb van der Meer Introduction 23 Pre-F€orster 23 Bottom Line 25 9000-Form, 9-Form, and Practical Expressions of the R06 Equation 26 Overlap Integral 28 Zones 31 Transfer Mechanisms 33 Kappa-Squared Basics 34 Ideal Dipole Approximation 35 Resonance as an All-or-Nothing Effect 36 Details About the All-or-Nothing Approximation of Resonance 39 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License jV j Contents 3.12 3.13 3.14 3.15 3.16 3.17 3.18 3.19 4 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12 4.13 5 5.1 5.2 5.3 5.4 5.4.1 5.4.2 5.4.2.1 5.4.2.2 5.4.2.3 5.4.2.4 5.4.2.5 5.4.3 Classical Theory Completed 41 Oscillator Strength–Emission Spectrum Relation for the Donor 42 Oscillator Strength–Absorption Spectrum Relation for the Acceptor 43 Quantum Mechanical Theory 44 Transfer in a Random System 47 Details for Transfer in a Random System 48 Concentration Depolarization 51 FRET Theory 1965–2012 52 References 59 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements 63 B. Wieb van der Meer, Daniel M. van der Meer, and Steven S. Vogel Two-Thirds or Not Two-Thirds? 63 Relevant Questions 65 How to Visualize Kappa-Squared? 65 Kappa-Squared Can Be Measured in At Least One Case 68 Averaging Regimes 70 Dynamic Averaging Regime 72 What Is the Most Probable Value for Kappa-Squared in the Dynamic Regime? 76 Optimistic, Conservative, and Practical Approaches 83 Comparison with Experimental Results 85 Smart Simulations Are Superior 90 Static Kappa-Squared 92 Beyond Regimes 101 Conclusions 102 References 103 How to Apply FRET: From Experimental Design to Data Analysis 105 Niko Hildebrandt Introduction: FRET – More Than a Four-Letter Word! 105 FRET: Let’s get started! 106 FRET: The Basic Concept 107 FRET: Inevitable Mathematics 112 F€orster Distance (or F€orster Radius) 112 FRET Efficiency 113 Determination by Donor Quenching 113 Determination by Acceptor Sensitization 113 Determination by Donor Quenching and Acceptor Sensitization 114 Determination by Donor Photobleaching 115 Determination by Acceptor Photobleaching 115 FRET with Multiple Donors and/or Acceptors 116 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License VI 5.5 5.5.1 5.5.2 5.5.3 5.5.3.1 5.5.3.2 5.5.3.3 5.6 5.6.1 5.6.1.1 5.6.2 5.6.3 5.6.4 5.6.4.1 5.6.4.2 5.6.4.3 5.6.4.4 5.7 6 6.1 6.2 6.3 6.3.1 6.3.2 6.3.3 6.3.4 6.3.5 6.3.6 6.3.7 6.4 6.4.1 6.4.2 6.4.3 6.4.4 6.4.5 6.4.6 6.4.7 FRET: The Experiment 118 The Donor–Acceptor FRET Pair 118 F€orster Distance Determination 119 The Main FRET Experiment 122 Steady-State FRET Measurements 123 Time-Resolved FRET Measurements 130 Interpretation of Time-Resolved FRET Data 133 FRET beyond F€orster 139 Time-Resolved FRET with Lanthanide-Based Donors 140 Terbium to Quantum Dot FRET Using Time-Resolved Donor Quenching and Acceptor Sensitization Analysis 141 BRET and CRET 147 Energy Transfer to Metal Nanoparticles (FRET, NSET, DMPET, NPILM, etc.) 148 Other Transfer Mechanisms 150 Electron Exchange Energy Transfer (Dexter Transfer) 151 Charge Transfer (Marcus Theory) 152 Plasmon Coupling 153 Singlet Oxygen Diffusion 154 Summary and Outlook 155 References 156 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 165 Kim E. Sapsford, Bridget Wildt, Angela Mariani, Andrew B. Yeatts, and Igor Medintz Introduction 165 Bioconjugation 166 Organic Materials 171 Ultraviolet, Visible, and Near-Infrared Emitting Dyes 171 Quencher Molecules 173 Environmentally Sensitive Fluorophores 175 Dye-Modified Microspheres/Nanomaterials 179 Dendrimers and Polymer Macromolecules 180 Photochromic Dyes 182 Carbon Nanomaterials 186 Biological Materials 188 Natural Fluorophores 188 Nonnatural Amino Acids 190 Green Fluorescent Protein and Derivatives 192 Light-Harvesting Proteins 200 DNA-Based Macrostructures/Nanotechnology 201 Enzyme-Generated Bioluminescence 201 Enzyme-Generated Chemiluminescence 209 jVII Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Contents j Contents 6.5 6.5.1 6.5.2 6.5.3 6.5.4 6.5.5 6.6 6.7 Inorganic Materials 211 Luminescent Lanthanide Complexes and Doped Nano-/ Microparticles 212 Luminescent Transition Metal Complexes 217 Noble Metal Nanomaterials (Gold, Silver, and Copper) 219 Silicon-Based Materials 222 Semiconductor Nanocrystals 223 Multi-FRET Systems 231 Summary and Outlook 236 References 236 Part Two Common FRET Techniques/Applications 269 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 271 Samantha Spindel, Jessica Granek, and Kim E. Sapsford Introduction 271 Small Organic Molecules and Synthetic Organic Polymers 272 Carbohydrate–Lipid 273 The Biotin–Avidin Interaction 273 Proteins and Peptides 275 Binding Proteins 275 Antigens and Epitope-Based Peptide Sequences 277 Peptide Sequences for Enzymatic Sensing 279 Antibodies 282 Nucleic Acid (DNA/RNA) 287 Molecular Beacons 288 Polymerase Chain Reaction and FRET 289 FRET Hybridization Probes 290 TaqMan 291 Scorpion Assay 292 Others 294 Isothermal Amplification Reactions and FRET 294 Clinical Applications of Nucleic Acid Detection Using FRET 295 Detection of Pathogens 295 Prognostic and Diagnostic Applications 296 Pharmacogenomics and Personalized Medicine 298 Aptamers 299 High-Throughput and Point-of-Care Devices 302 PoC Technology Advances 302 PoC Material Advances 304 Conclusions 305 References 305 7.1 7.2 7.3 7.4 7.5 7.5.1 7.5.2 7.5.3 7.6 7.7 7.7.1 7.7.2 7.7.2.1 7.7.2.2 7.7.2.3 7.7.2.4 7.7.3 7.7.4 7.7.4.1 7.7.4.2 7.7.4.3 7.8 7.9 7.9.1 7.9.2 7.10 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License VIII 8 8.1 8.2 8.2.1 8.2.1.1 8.2.1.2 8.2.1.3 8.2.2 8.2.3 8.2.4 8.3 8.3.1 8.3.2 8.3.3 8.3.3.1 8.3.3.2 8.4 8.4.1 8.4.2 8.5 9 9.1 9.2 9.3 9.3.1 9.3.2 9.4 9.4.1 9.4.2 9.5 9.6 9.6.1 9.6.2 9.7 9.7.1 9.7.2 9.8 9.9 Single-Molecule Applications 323 Thomas Pons Introduction 323 Single-Molecule FRET of Immobilized Molecules 324 Experimental Setup 324 Molecule Immobilization 324 Fluorophore Photostability 325 Optical Setup 326 Data Analysis 326 Applications 329 Analyzing Complex FRET Trajectories 334 Single-Molecule FRET of Freely Diffusing Molecules 336 Experimental Setup 336 Applications 337 Advanced Solution smFRET Methods 343 Alternate Laser Excitation 343 Multiparameter Fluorescence Detection 344 Single-Molecule FRET Studies Involving Multiple FRET Partners 346 Multistep FRET 347 Multi-Acceptor and Multi-Donor Systems 348 Conclusions and Perspectives 351 References 353 Implementation of FRET Technologies for Studying the Folding and Conformational Changes in Biological Structures 357 Philip J. Robinson and Cheryl A. Woolhead Introduction to Using FRET in Biological Systems 357 F€orster Formalism in the Determination of Biological Structures 358 FRET Experiments in Complex Biological Systems 360 The Importance of Experimental Design 360 Site-Specific Labeling and Choosing the Most Effective FRET Pair 361 Biological Model System 1: The Ribosome 362 Intersubunit Rotation within the Ribosome 363 Dynamic Intrasubunit Movement Within the Ribosome 365 Biological System 2: Nascent Polypeptide Structure 365 Biological System 3: Chaperone-Mediated Protein Folding 368 Signal Recognition Particle 368 Trigger Factor 369 Biological System 4: Mature Protein Folding Intermediates 371 Unfolding Kinetics of Monellin 372 Intermediate Folding State of the Src Homology 3 Domain 374 Biological System 5: Intersubunit Distance in Multimeric Protein Complexes 375 Biological System 6: Protein–Protein Interactions in the Assembly of Protein Polymers 378 jIX Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Contents 9.9.1 9.9.2 9.10 9.10.1 9.10.2 9.10.3 10 10.1 10.2 10.2.1 10.2.2 10.3 10.3.1 10.3.2 10.4 10.4.1 10.4.2 10.4.3 10.4.4 10.5 Part Three 11 11.1 11.1.1 11.1.1.1 11.1.1.2 11.1.1.3 11.1.2 11.1.2.1 11.1.2.2 11.1.2.3 11.1.2.4 11.1.2.5 FtsZ Assembly and Subunit Exchange 379 Defining the Molecular Link in Serpin Polymers 380 Biological System 7: FRET in Nucleic Acid Systems 385 Determining the Structure and Configuration of DNA Junctions 386 Measuring the Opening and Closing of a Nanoscale DNA Box 388 FRET Between a DNA Polymerase and Its Substrate 390 References 392 FRET-Based Cellular Sensing with Genetically Encoded Fluorescent Indicators 397 Jonathan C. Claussen, Niko Hildebrandt, and Igor Medintz Introduction 397 Enzymes 399 Kinase Activity/Phosphorylation 399 Protease Activity 403 Metabolites 407 Sugars 407 Glutamate 410 Second Messengers 412 cAMP 412 cGMP 415 Nitric Oxide 417 Calcium 419 Conclusions 421 References 423 FRET with Recently Developed Materials 431 FRET with Fluorescent Proteins 433 Hiofan Hoi, Yidan Ding, and Robert E. Campbell Introduction to FPs 433 Wild-Type FPs 433 Natural Sources 433 Structure 434 Chromophore Formation 436 Engineered FPs for FRET Applications 438 Overview 438 Blue–Green FRET Pairs 440 Cyan–Yellow FRET Pairs 441 FRET with Orange, Red, and Far-Red FPs 443 Atypical FPs Useful for FRET Applications 445 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j Contents X 11.1.3 11.2 11.2.1 11.2.1.1 11.2.1.2 11.2.1.3 11.2.2 11.2.2.1 11.2.2.2 11.2.2.3 11.2.3 11.2.3.1 11.2.3.2 11.2.3.3 11.2.3.4 11.2.4 11.2.4.1 11.2.4.2 11.2.4.3 11.3 Why Use FPs for FRET? 446 Using FPs for FRET Imaging 446 Photophysical Properties and Typical F€ orster Radii 446 Overview 446 Spectral Overlap 447 Orientation Factors 449 Potential Sources of Artifacts During FRET Imaging 450 Photobleaching 450 Photoconversion 451 pH Dependence 452 Biochemical and Structural Considerations 453 General Considerations when Labeling Proteins with FPs 453 Labeling Proteins for Intermolecular FRET Experiments 454 Labeling Proteins for Intramolecular FRET Experiments 454 FP Oligomerization and FRET Efficiency 455 Applications and Examples 458 Overview 458 FRET Biosensor Case Study 459 FRET between FPs and Other Donor or Acceptor Materials 460 Conclusions 462 References 463 12 Semiconductor Quantum Dots and FRET 475 W. Russ Algar, Melissa Massey, and Ulrich J. Krull Introduction 475 A Quick Review of FRET 476 Quantum Dots 477 A Brief History 478 The Structure of Quantum Dots: The Core 478 The Optical Properties of Quantum Dots 480 Overcoming the Limitations of Molecular Fluorophores 482 The Structure of Quantum Dots: The Shell 483 Quantum Confinement 485 Quantum Dot Photophysics 488 Quantum Dot Synthesis 491 Quantum Dot Coatings 493 Quantum Dot Bioconjugation 496 Quantum Dot Nomenclature in This Chapter 499 Quantum Dots and FRET 499 Quantum Dots as Donors 499 Applicability of the F€orster Formalism 502 QDs as Acceptors 504 The Importance of Bioconjugate Chemistry 506 Quantum Dots as Donors in Biological Applications 508 Association and Dissociation to Modulate QD-FRET 508 12.1 12.2 12.3 12.3.1 12.3.2 12.3.3 12.3.4 12.3.5 12.3.6 12.3.7 12.3.8 12.3.9 12.3.10 12.3.11 12.4 12.4.1 12.4.2 12.4.3 12.4.4 12.5 12.5.1 jXI Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Contents j Contents 12.5.1.1 12.5.1.2 12.5.1.3 12.5.1.4 12.5.1.5 12.5.1.6 12.5.2 12.5.3 12.5.4 12.5.5 12.5.6 12.5.6.1 12.5.6.2 12.5.6.3 12.5.6.4 12.5.6.5 12.5.7 12.6 12.6.1 12.6.2 12.6.3 12.6.4 12.7 12.7.1 12.7.2 12.7.2.1 12.7.2.2 12.8 12.8.1 12.8.2 12.9 13 13.1 13.2 13.2.1 13.2.2 13.2.3 13.3 13.4 13.4.1 13.4.2 Bioanalysis of Carbohydrates 509 Homogeneous Immunoassays 510 Hybridization Assays 511 Bioanalyses Using Aptamers and DNAzymes 516 Bioanalysis of Hydrolytic Enzymes 519 Gene Delivery 524 Changes in Distance to Modulate QD-FRET 524 Conformational Insights from QD-FRET 528 Dynamic Modulation of the Spectral Overlap Integral and QD-FRET 530 Single-Pair QD-FRET 534 Solid-Phase QD-FRET 540 Biomolecular Surface Tethers 542 Chemical Conjugation to an Interface 544 Interfacial Ligand Exchange 545 Electrostatic Immobilization 547 Advantages of Immobilized QDs 548 Photodynamic Therapy 549 Quantum Dots as Acceptors in Biological Applications 552 Chemiluminescence Resonance Energy Transfer (CRET) 553 Bioluminescence Resonance Energy Transfer (BRET) 555 Lanthanide Donors 560 Quantum Dot Donors (for Quantum Dot Acceptors) 565 Energy Transfer between Quantum Dots and Other Nanomaterials 569 Gold Nanoparticles 569 Carbon Nanomaterials 575 Graphene and Graphene Oxide 575 Carbon Nanotubes 577 Nonbiological Applications of Quantum Dots and FRET 578 Photovoltaic Cells 580 Light-Emitting Diodes (LEDs) 582 Summary 583 References 584 Multistep FRET and Nanotechnology 607 Bo Albinsson and Jonas K. Hannestad Introduction 607 Fundamentals of Multistep FRET 608 Hetero-FRET 609 Multicolor FRET and Alternating-Laser Excitation 611 Homo-FRET 612 Energy Transfer in Photosynthesis 615 Photonic Wires and Multistep FRET in Nanotechnology 617 Photonic Wires 617 Beyond Wires 628 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License XII 13.4.3 13.4.4 13.4.5 13.4.6 13.5 13.6 Light Harvesting 632 Functional Control 638 Quantum Dots in Multistep FRET 641 Potential Outputs and Uses for Channeled Excitation Energy 643 Summary 647 Note Added in Proof 648 References 648 Part Four Supporting Information and Conclusions 655 14 Data 657 Alice G. Byrne, Matthew M. Byrne, George Coker III, Kelly B. Gemmill, Christopher Spillmann, Igor Medintz, Seth L. Sloan, and B. Wieb van der Meer Tables before 1987 658 Introduction to the Table of Traditional Chromophores 658 F€orster Distances and Other FRET Data before 1994 703 F€orster Distances for Traditional Probes More Recent Than 1993 703 FRET Data on Fluorescent Proteins 703 FRET Data on Quantum Dots 742 Donor–Acceptor Pairs with a F€orster Distance in a Given Range 742 Table–Reference Directory 744 References 745 14.1 14.2 14.3 14.4 14.5 14.6 14.7 14.8 15 15.1 15.2 15.3 15.4 15.5 15.6 15.7 Outlook on FRET: The Future of Resonance Energy Transfer 757 A Rosy Crystal Ball View of FRET 757 Thomas M. Jovin Do Not Ask What FRET Can Do for You, Ask What You Can Do for FRET 757 B. Wieb van der Meer FRET: Future Research with an Exciting Technology 758 Niko Hildebrandt Future of FRET 760 Kim E. Sapsford Outlook on Single-Molecule FRET 760 Thomas Pons Outlook on FRET with fluorescent proteins 761 Robert E. Campbell Luminescent Nanoparticles: Scaffolds for Assembling “Smarter” FRET Probes 762 W. Russ Algar References 764 Index 767 jXIII Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Contents Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Preface Several factors over the last few years have come together to contribute to the origin and development of this book. First and foremost is the incredible expansion in the application of FRET and its derivative techniques, especially in the biological sciences. Web of Science (www.thomsonreuters.com) provides more than 12,000 items using the search term FRET as a unique topic, while PubMed (www.ncbi.nlm .nih.gov/pubmed) provides almost 6000 items. The latter site also allows tracking of these items over time with 0 hits shown for 1979 and >1000 predicted for 2013 – an impressive and eye-opening increase. This stands in stark contrast to the years 1970–1990 that have less than 20 publications in total. This is not to say that nothing significant happened during this time period, but rather it reflects how specialized the field was and reminding us also of how poor the performance of journal citation and referencing tools were before the 1990s. In terms of just citations alone, F€orster’s original 1948 paper in Annalen der Physik has to its credit a remarkable 5000 citations, although it is safe to say that only a minority of those who cite this article have read it (especially in the original German). Indeed, some scientists consider this to be one of the most cited papers that has never been actually read. Development of and access to a wide range of versatile fluorescent materials in conjunction with improved, easy-to-use and yet incredibly sophisticated microscopes and fluorometers have coincided with, helped drive, and also increased FRET usage. Fluorophores that are utilized in FRET now commonly encompass organic dyes, fluorescent proteins, semiconductor quantum dots, metal chelates, various noble metal and other nanoparticles, intrinsically fluorescent amino acids, biological cofactors, and polymers, to name but a few members of this growing library. Hand in hand with materials development is the growing availability of numerous reactive and bioorthogonal chemistries to site–specifically attach such fluorophores to all types of biological molecules ranging from proteins to DNA. This, in conjunction with FRETs unique ability to consistently provide nanoscale inter- and intramolecular separation distances, has meant that its utility is also rapidly growing in structural studies of biomolecules and biological complexes. We have also seen the implementation of intracellularly expressed fluorescent protein-based FRET sensors expand so rapidly over the past 15 years that it is now not uncommon to encounter students who do not know where this technology originated from (Roger Tsien, University of California San Diego). Perhaps this is the ultimate form of a Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License jXV j Preface compliment in science – when something becomes so commonplace and widely accepted that people forget who invented it. Concomitant with this rapid development of materials, commercial microscope systems combined with sophisticated analytical software are now widely available providing direct access to many different FRET techniques and their derivatives. Another proof that FRET techniques have come of age is the recent meeting entitled “F€ orster Resonance Energy Transfer in the Life Sciences” held at the Max Planck Institute for Biophysical Chemistry in G€ottingen, Germany, in March 2011. This relaxed and wonderfully stimulating discussion meeting organized by Donna Arndt-Jovin, Stefan Hell, Thomas Jovin, Claus Seidel, and J€ urgen Troe focused on all the different aspects of FRET from analytical techniques and microscopy to new materials. That FRET could be the stand-alone subject of an international scientific meeting speaks volumes to its growing utility. In addition, not many people realize that the genomics revolution of the past 15–20 years owes a large debt to the use of FRET. Richard Mathies and Alex Glazer at the University of California Berkeley were among the first to realize that use of a dye-based FRET system could drastically simplify the instrumental requirements for DNA sequencing. By linking a single donor dye with four different acceptors (representing the four DNA bases), they were able to provide a set of four common DNA primers. These constructs used FRET to create four spectrally wellseparated windows that could be excited by a single laser wavelength in any electrophoretic system instead of requiring two or three separate lasers. This strategy quickly became the workhorse of DNA analysis and, as is typical for any successful and proprietary technology, this also became the subject of considerable litigation. Similar FRET systems form the basis for numerous genotyping tests such as the Taqman assay that have also contributed quite considerably to genomics. A variety of other FRET assay formats for monitoring enzymatic activity and the like have also become quite commonplace in biosensing, biological research, and drug discovery. Niko and I both come from laboratories that are very interested in understanding how newer materials such as quantum dots and/or long-lifetime rare-earth chelates engage in FRET and other forms of energy transfer. These materials provide for fascinating energy transfer configurations that were not described or even considered in F€orster’s seminal treatise. For example, can FRET occur when the acceptor is as well excited as a donor while manifesting a much longer excited state lifetime as in the case of pairing an organic dye donor such as fluorescein with a red-emitting quantum dot. With questions such as this in mind, one of the main FRET resources that is almost invariably consulted first in the pursuit of appropriate background is Lakowicz’s excellent primer – Principles of Fluorescence Spectroscopy (Springer). Although still one of the most readable textbooks ever published, this resource provides only a limited amount of data and discussion on the intricacies of FRET. Van Der Meer’s Resonance Energy Transfer Theory and Data (Wiley-VCH Verlag GmbH) is far more detailed about FRET mechanics and is another well-cited reference in this area; however, this has unfortunately fallen out of print and is quite hard to find. Thus, it was that we both found ourselves lamenting the lack of an up-to-date and detailed resource/primer on all aspects of FRET when Wiley-VCH Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License XVI graciously approached us about undertaking a book on an important scientific area of our choice. It did not take us long to decide the subject that we were going to propose. The authors who contributed the individual chapters in this book reflect not only some of the principal experts in the field but also the wide diversity of FRET application itself. We very much appreciate the support of leaders in this field, including Robert E. Campbell, Bo Albinsson, and Jonas Hannestad. We were also overjoyed when Wieb van Der Meer agreed to make several major contributions to this project. The book is divided into four major parts. Part One provides some background, theory, and key concepts. This part begins with a personal remembrance of Theodor F€orster by one of his former students and colleagues Herbert Dreeskamp (Chapter 1). Thomas Jovin then notes the recent passing of Robert Clegg and Elizabeth Jares-Erijman by describing their important contributions to FRET (Chapter 2). Wieb van der Meer then takes us in detail through the F€ orster theory and tackles the always important yet continually vexing issue of kappa-squared (Chapters 3 and 4). Niko Hildebrandt then provides a detailed primer on how to apply FRET – from experimental design to data analysis (Chapter 5). Finally, Kim Sapsford updates a previous 2006 paper that describes the ever-growing FRET toolbox of diverse fluorophores (Chapter 6). Part Two of the book focuses on some common FRET techniques and applications. Kim Sapsford and her colleagues from the U.S. Food and Drug Administration again contribute with a discussion of FRET application for in vitro sensing and diagnostics (Chapter 7). Thomas Pons then reviews single-molecule FRET applications, which represent another rapidly growing and important area (Chapter 8). Cheryl Woolhead provides a description of FRET utility in the determination of protein, peptide, and other biological structures (Chapter 9). This part ends with a contribution from Jonathan Claussen on FRETbased cellular biosensing with genetically encoded fluorescent indicators (Chapter 10). Part Three looks at recent developments starting with the use of fluorescent proteins from Robert Campbell (Chapter 11). This is followed by a review of FRET usage with semiconductor quantum dots from W. Russ Algar and colleagues (Chapter 12) along with an overview of the growing area of multistep FRET from Bo Albinsson and Jonas Hannestad (Chapter 13). The concluding Part Four includes a detailed and vastly updated series of supporting tables on FRET pairs and F€orster distances collected and collated by Wieb van der Meer (Chapter 14). These tables were so useful in his previous book that we could not let this opportunity to update them go unused. Finally, some of the authors provide their own outlook on and perspectives of FRET (Chapter 15). We want to thank all of the authors for not only their time and contributions but also for their incredible patience with us as this book slowly came together. The same is true for our coworkers at Wiley, including Eva-Stina M€ uller and Heike N€ othe. We have tried to focus on the important aspects that will both help the FRET novice and reinforce the understanding of a seasoned FRET user. Given the details and growth of this technology, we realize that we could not include everything we wanted and our apologies are further extended for any and all omissions along with any errors. We admit that we were rather naïve and a little overly hopeful in some of our initial jXVII Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Preface j Preface ideas and desires for this book. A detailed historical accounting of F€ orster’s life and work is still missing from the literature and although we tried repeatedly, we could not bring this to fruition for the current book. We also wanted a dedicated Web site to accompany this book where FRET data, especially on newer donor–acceptor pairs, could be continuously updated along with providing a forum for discussion and solicitation of experimental advice. Alas, such a home is not to be found as of yet. Finally, although he tried quite valiantly, Niko could not convince any television producers or comic book publishers to introduce a new superhero for children (and scientists), namely, Captain FRET who solves complicated situations with the application of resonance energy transfer while carefully explaining the subsequent photophysical analysis for the layman. If there is ever an update to this book, we will redouble our efforts to bring these additional ideas to reality. In the preface to Resonance Energy Transfer Theory and Data, Wieb van Der Meer outlines how most scientists with interest in energy transfer can be subdivided into two groups: those interested in homotransfer, the homotransferites (primarily biochemists), and those interested in heterotransfer, the heterotransferites (primarily physical chemists). In the same manner as him, this book is written for all and not just one group. However, in adhering to the culture of our times and rather than differentiating between groups, we would like to add a new all-inclusive description to the FRET user anthology; if you are able to utilize FRET successfully (in any form), then you should be considered and referred to as a “FRET jock” and this should be a moniker of distinction and pride among your scientific colleagues. It is our fervent hope that well-worn copies of this book find their way onto your office and laboratory shelves. Niko Hildebrandt Igor L. Medintz Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License XVIII List of Contributors Bo Albinsson Chalmers University of Technology Department of Chemistry and Biotechnology/Physical Chemistry 41296 Gothenburg Sweden W. Russ Algar University of British Columbia Department of Chemistry 2036 Main Mall Vancouver British Columbia V6T 1Z1 Canada Alice G. Byrne Western Kentucky University Department of Physics and Astronomy 1906 College Heights Blvd Bowling Green KY 42101 USA Matthew M. Byrne Western Kentucky University Department of Physics and Astronomy 1906 College Heights Blvd Bowling Green KY 42101 USA Robert E. Campbell University of Alberta Department of Chemistry 11227 Saskatchewan Drive Edmonton Alberta T6G 2G2 Canada Jonathan C. Claussen U.S. Naval Research Laboratory Center for Bio/Molecular Science and Engineering Code 6900 4555 Overlook Avenue, SW Washington DC 20375 USA and George Mason University College of Science Fairfax VA 22030 USA Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License jXIX George Coker, III Western Kentucky University Department of Physics and Astronomy 1906 College Heights Blvd Bowling Green KY 42101 USA Yidan Ding University of Alberta Department of Chemistry 11227 Saskatchewan Drive Edmonton Alberta T6G 2G2 Canada Herbert Dreeskamp Technische Universit€at Braunschweig Germany Kelly B. Gemmill U.S. Naval Research Laboratory Center for Bio/Molecular Science and Engineering Code 6900 4555 Overlook Avenue, SW Washington DC 20375 USA Jessica Granek U.S. Food and Drug Administration CDRH/OSEL/DB WO64 RM3028 HFZ-110 10903 New Hampshire Avenue Silver Spring MD 20993 USA Jonas K. Hannestad Chalmers University of Technology Department of Chemistry and Biotechnology/Physical Chemistry 41296 Gothenburg Sweden Niko Hildebrandt Universite Paris-Sud Institut d’Electronique Fondamentale NanoBioPhotonics Bâtiment 220 91405 Orsay Cedex France Hiofan Hoi University of Alberta Department of Chemistry 11227 Saskatchewan Drive Edmonton Alberta T6G 2G2 Canada Thomas M. Jovin Max Planck Institute for Biophysical Chemistry Laboratory of Cellular Dynamics 37077 G€ ottingen Germany Ulrich J. Krull University of Toronto Mississauga Department of Chemical and Physical Sciences 3359 Mississauga Rd. North Mississauga Ontario L5L 1C6 Canada Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j List of Contributors XX Angela Mariani U.S. Food and Drug Administration CDRH/OSEL/DB WO64 RM3028 HFZ-110 10903 New Hampshire Avenue Silver Spring MD 20993 USA Melissa Massey University of Toronto Mississauga Department of Chemical and Physical Sciences 3359 Mississauga Rd. North Mississauga Ontario L5L 1C6 Canada Igor Medintz U.S. Naval Research Laboratory Center for Bio/Molecular Science and Engineering Code 6900 4555 Overlook Avenue, SW Washington DC 20375 USA Thomas Pons ESPCI–CNRS–UPMC (UMR8213) Laboratoire de Physique et d’Etude des Materiaux 10, rue Vauquelin 75005 Paris France Philip J. Robinson University of Glasgow College of Medical, Veterinary and Life Sciences Institute of Molecular, Cell and Systems Biology Glasgow Lanarkshire G12 8QQ Great Britain Kim E. Sapsford U.S. Food and Drug Administration CDRH/OSEL/DB WO64 RM3028 HFZ-110 10903 New Hampshire Avenue Silver Spring MD 20993 USA Seth L. Sloan Western Kentucky University Department of Physics and Astronomy 1906 College Heights Blvd Bowling Green KY 42101 USA Christopher Spillmann U.S. Naval Research Laboratory Center for Bio/Molecular Science and Engineering Code 6900 4555 Overlook Avenue, SW Washington DC 20375 USA jXXI Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License List of Contributors j List of Contributors Samantha Spindel U.S. Food and Drug Administration CDRH/OSEL/DB WO64 RM3028 HFZ-110 10903 New Hampshire Avenue Silver Spring MD 20993 USA Bridget Wildt U.S. Food and Drug Administration CDRH/OSEL/DB WO64 RM3028 HFZ-110 10903 New Hampshire Avenue Silver Spring MD 20993 USA B. Wieb van der Meer Western Kentucky University Department of Physics and Astronomy 1906 College Heights Blvd Bowling Green KY 42101 USA Cheryl A. Woolhead University of Glasgow College of Medical, Veterinary and Life Sciences Institute of Molecular, Cell and Systems Biology Glasgow Lanarkshire G12 8QQ Great Britain Daniel M. van der Meer TelaPoint 9500 Ormsby Station Road, Suite 402 Louisville KY 40223 USA Steven S. Vogel National Institutes of Health National Institute on Alcohol Abuse and Alcoholism Laboratory of Molecular Physiology 5625 Fishers Lane Bethesda MD 20892 USA Andrew B. Yeatts U.S. Food and Drug Administration CDRH/OSEL/DB WO64 RM3028 HFZ-110 10903 New Hampshire Avenue Silver Spring MD 20993 USA Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License XXII Part One Background, Theory, and Concepts FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j1 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 1 How I Remember Theodor F€ orster) Herbert Dreeskamp I first met Theodor F€orster in 1959, after my postdoctoral years in the United States, at a conference on fast reactions organized by Manfred Eigen in the Harz Mountains. Very prominent scientists attended, such as Eyring, Noyes, and the three men seen in Figure 1.1 (from left to right) George Porter, Theodor F€ orster, and Albert Weller. Half a century ago “fast reaction” meant flash photolysis in the microsecond range by Norrish and Porter or relaxation methods by Eigen. As I remember, it was F€ orster who pointed out clearly that the term “fast” characterized our technical facilities at that time rather than the scientific problem at hand. The time range of fast chemical reactions may better be characterized by the rearrangement of electrons in the 1016 s range, the vibration of nuclei in the 1012 s range, or the deactivation of electronically excited states in the 109 s range. Thus, if there are reactions of electronically excited states – and after all, molecules do have characteristically different properties in their different electronic states – you will be able to investigate these reactions in the nanosecond range by just studying fluorescence, which is emitted in competition to these photochemical reactions. And since F€ orster had, at the beginning of his career, studied the absorption spectra of organic compounds, that is, the electronic structure of their ground and excited states, he was able to find the proteolytic reactions of aromatic compounds as the classical example of using fluorescence to investigate fast chemical reactions. To me this example shows clearly, in a nutshell, F€orster’s approach to the scientific problem and why he was so extremely successful in opening new avenues in photochemistry. Sometimes it was said that he was gifted by a remarkable intuition. I am sure his intuition was the result of strict devotion to science, very hard work, his enormous ) This chapter is based on a talk given by the author at the International Bunsen Discussion Meeting on “Light Harvesting and Solar Energy Conversion,” March 29, 2010, Stuttgart-Hohenheim, commemorating the 100th birthday of Theodor F€ orster (1910–1974). The author studied physics in Bonn and Paris, spent decisive years 1960–1970 with F€orster in Stuttgart, and was professor of physical chemistry at the Technische Universit€at Braunschweig. He thanks Dr. Eberhard F€orster, son of Theodor F€orster, for the pictures used in this chapter. FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j3 j 1 How I Remember Theodor F€orster Figure 1.1 knowledge of the literature, and his insistence to reach a complete understanding of the problem at hand. Theodor F€orster was a son of Frankfurt am Main, like Otto Hahn (as shown in Figure 1.2, second from left, and F€orster is the first from right). He got a training there as a theoretical physicist when both he and quantum mechanics were quite young (Figure 1.3). As an assistant to Karl Friedrich Bonhoeffer in Leipzig, he came under the influence of such eminent men – besides Bonhoeffer of course – as Heisenberg, Kautsky, and, I think particularly, Peter Debye. Since those Leipzig days there is the most remarkable and efficient interplay between theory and experiment in the work of Theodor F€orster. Figure 1.2 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4 Figure 1.3 His aim was to search for the most appropriate solution of the scientific problem, or in his favorite words: “Die richtige Deutung einer Beobachtung” (The correct interpretation of an observation). This included taking carefully all information into account, separating the important from the trivial, designing a simple experiment, and arriving at the correct interpretation, if possible, without any too elaborate computer analysis. For me, this picture (Figure 1.4) from the Posen or G€ ottingen years – the 1940s – may illustrate what I tried to say: Brains seem to be more important than fancy equipment or powerful computers. Very often it was both a relief and a delight for all of us who were present, when after a somewhat incomprehensible seminar talk he would stand up and quite politely say, “If I understand you correctly, you meant to say this and that . . . ” and he would give in a few words a lucid interpretation of the topic at hand. I once asked F€orster how to grade a thesis paper, and he advised me to be not too strict. But for him, Theodor F€orster, his scientific work had to meet the highest standards. Things had to be correct, of course, but equally important: it had to include the most concise analysis of the problem, a perfect logic of the solution, and a clear statement on the significance of the results. He did not publish much, but the things he did publish can be a source of inspiration still today. j5 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 1 How I Remember Theodor F€orster j 1 How I Remember Theodor F€orster Figure 1.4 You may know that the phenomenon of fluorescence depolarization was an important step that ultimately led to an understanding of electronic excitation energy transfer between molecules. After a large amount of empirical material had been accumulated by others, F€orster, in a lucidly written review article, gave a brilliant analysis of this effect and brought a long discussion ultimately to an end. Thus, may I advice you, once in a while, to take your time off from the lab and go to the library and study a paper of his, or better still his most admirable monograph “Fluoreszenz organischer Verbindungen.” You may be rewarded by getting a hint on how problems may be solved by putting them in the right perspective, a strategy in which Theodor F€orster was a superb master. My picture of F€orster would be incomplete without remembering how much he enjoyed the company of colleagues, or of his students, for example, at a Christmas party in the lab (Figure 1.5). Very often prominent colleagues from abroad came to Stuttgart, gave a talk, and certainly were invited by him and his wife Martha (Figure 1.6) to their home. Regularly, younger members of the department were also invited to these evenings. For me, certainly the most memorable of these meetings was when James Franck visited Stuttgart, I think in 1964 (Figure 1.6). You all will know the fundamental work in atomic physics done in Berlin and G€ottingen by Franck and Hertz in the 1920s, or the direct proof of a radiationless energy transfer between atomic systems by Franck Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6 Figure 1.5 and Cario. But I think we should also remember that James Franck was the initiator of the Franck Report of 1945. In the last picture (Figure 1.7), you see James Franck and Theodor F€ orster many years after the war, evidently discussing at a scientific conference. Also the topic of their discussion was – I am pretty sure, – the phenomenon of light harvesting, energy transfer, and photosynthesis, questions that fascinated both these men for many years. Franck gave the first experimental proof that the electronic energy may be exchanged radiationlessly among atoms, and F€ orster, some 25 years later, on the basis of his deep understanding of quantum mechanics, gave us the theory of the Figure 1.6 j7 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 1 How I Remember Theodor F€orster j 1 How I Remember Theodor F€orster Figure 1.7 nontrivial transfer of electronic energy in molecular systems, the “F€ orster resonance energy transfer” (FRET), which gave us a formula that has become extremely important in biological sciences. The contributions of Theodor F€orster to modern photochemistry are most impressive, but equally fascinating to me is the way in which he elaborated these things. If you have a look at his strategy, I am certain you will have a good chance to profit also from this aspect of the work of Theodor F€ orster. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8 2 Remembering Robert Clegg and Elizabeth Jares-Erijman and Their Contributions to FRET Thomas M. Jovin Robert M. Clegg Elizabeth A. Jares-Erijman This is a rather personal account, yet not biographical, of a scientific “family” bound by circumstances and a common pervasive scientific theme. It is perhaps the essential nature of F€orster resonance energy transfer (FRET) – a near-field resonance phenomenon – that engenders “resonating” interactions between individuals. For Robert (“Bob”) Clegg and Elizabeth (“Eli”) Jares-Erijman, as well as for the redactor of this account, Thomas (“Tom”) Jovin, there were distinct circles of scientific and personal influences that dictated how FRET entered their lives and careers. As in the case of most scientists, the initial event was exposure to key literature. In the emerging FRET field after World War II, a number of highly cited original papers and reviews stimulated innumerable scientists to incorporate FRET into their conceptual and experimental strategies. The reason lay with their authors, leading protagonists and innovators, who in historical order included Theodor F€ orster [1], Gregorio Weber [2], Izchak Steinberg [3], Lubert Stryer [4], and Ludwig (Lenny) Brand [5]. There are other equally valuable sources; I cite these because they predominated in my case and were also highly influential for Bob and Eli, ultimately leading to their own valuable contributions. FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j9 j 2 Remembering Robert Clegg and Elizabeth Jares-Erijman and Their Contributions to FRET 2.1 Biographical Sketch of Bob Clegg Robert MacDonald Clegg succumbed to cancer on October 15, 2012. His tragic death represents an immense loss not only to the members of his immediate family (wife Margitta, and sons Benjamin, Niels, and Robert), but also to his very extensive “scientific family” of colleagues and friends. Bob received his PhD in physical chemistry from Cornell University in 1974, having worked with Elliot Elson on the theory and practice of rapid chemical kinetics, specifically chemical relaxation following pressure perturbation. Elliot was a product of the lab of Buzz Baldwin in the Stanford Biochemistry Department and his influence on Bob’s view of science and research cannot be overemphasized. As a postdoc (with me) in the Department of Molecular Biology at the Max Planck Institute for Biophysical Chemistry, Bob quickly demonstrated his depth of knowledge and unique leadership and innovative skills. He assumed the position of Senior Staff Research Associate with an independent group in 1976. Over the next two decades, he pursued numerous lines of research, devising and applying quantitative thermodynamic, kinetic, and spectroscopic techniques, particularly fluorescence – which he had not used previously – to studies of macromolecular systems such as RNA polymerase. He became one of the best expounders worldwide of the theory and practice of FRET (energy transfer) and was involved in pioneering implementations and applications of fluorescence lifetime imaging microscopy (FLIM). Bob is well remembered in G€ ottingen for being “big and broad” in both science and physique, but also for his invariably cheerful and gentle disposition. He was ever ready to offer help in the form of advice or action. He was everybody’s friend. During a sabbatical leave in 1996 at the University of Illinois (UIC) in Champaign, Bob established a close working relationship with fluorescence pioneer Gregorio Weber. Weber’s distinguished disciple Enrico Gratton induced Bob’s repatriation to the United States (UIC) in 1998, with an appointment as professor in the Departments of Physics and Bioengineering and as a member of the faculty of the Biophysics program (at the time of his death, Bob was its director and an affiliate of the Institute for Genomic Biology). In this academic environment, Bob quickly established himself as a leading researcher in numerous biophysical disciplines as well as an extraordinarily dedicated and capable teacher. Bob was both an excellent experimentalist and theoretician, and consistently sought a fundamental understanding of the phenomena under investigation. In so doing, he displayed a unique capacity for deciphering structure–function relationships in complex systems involving transitions in molecular conformation and association of proteins and nucleic acids. His contributions to the expansion of optical microscopy into new fields of biology and biotechnology were also numerous and profound. The breadth of Bob’s interests and associations was reflected in his membership in the Biophysical Society, the American Physical Society, FASEB, the Optical Society of America, and the American Chemical Society. In 2009, the Biophysical Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 10 Society recognized his contributions to fluorescence with the Weber Award for Excellence in Theory and Experiments in Fluorescence, Fluorescence Subgroup. That same year, he received from the Society for Experimental Biology and Medicine the Alan MacDiarmid Best Paper Award in the interdisciplinary research category. The paper in question was “Engineering redox-sensitive linkers for genetically encoded FRET-based biosensors,” a typical (for Bob) synthesis of biology, chemistry, and physics. 2.2 Biographical Sketch of Eli Jares-Erijman Elizabeth Andreas Jares-Erijman, professor in the Department of Organic Chemistry at the University of Buenos Aires, Argentina, died of cancer on September 29, 2011. She was 50 years old. As in the case of Bob Clegg, we, the scientific community, lost an excellent, innovative scientist, a stimulating teacher, and a wonderful friend. A chemist by training, Elizabeth Eli received her PhD from the University of Buenos Aires in Argentina in 1989. After a postdoctoral period in the Department of Chemistry at UIC, she was transferred to my lab in G€ ottingen in 1993, accompanied by her husband Leonardo (Leo) – also a postdoc in the department – and her daughter Paula (see more details later). Three years later, Eli returned to Argentina and rejoined the Department of Organic Chemistry in the Faculty of Exact and Natural Sciences. She advanced through the academic hierarchy and occupied a pivotal role in the teaching and research activities of the department. Eli had her second child, Florencia, in 1998. In 2004, the Max Planck Institute for Biophysical Chemistry recognized her seminal contributions and involvement with the research program of the institute and proposed her for appointment as Head of a Max Planck Partner Group of the institute. This came to pass after an evaluation by an outside commission. Hers was the first partner group to be established in Argentina, in fact, the first in all of Latin America. Eli was one of the few individuals in Latin American science who crossed rigid departmental lines in order to establish a comprehensive and systematic research effort in what is currently designated as chemical and supramolecular biology. She established a Laboratory of Nanotools and Bioimaging to promote the design and use of novel organic probes and multifunctional nanoparticles as biosensors and “nanoactuators.” New implementations of FRET, for example, exploiting the phenomenon of photochromism, were an important feature in many of these systems. However, the biological applications were always at the research focus, as illustrated in recent publications devoted to a-synuclein, the “amyloid” protein in Parkinson’s disease (PD). Eli was an excellent citizen of her scientific community, serving in many commissions, both at the local and at the national levels. By all indicators, she was an inspired and very competent teacher. A significant indicator of her persuasive and inspiring leadership is the quality and success of the people who worked with her. She received numerous awards for her scientific achievements, which included j11 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 2.2 Biographical Sketch of Eli Jares-Erijman j 2 Remembering Robert Clegg and Elizabeth Jares-Erijman and Their Contributions to FRET the prize Eduardo G. Gros and the prize Bernado A. Houssay (an Argentine Nobel laureate), and was nominated in 2011 for the prestigious UNESCO-L’Oreal Award for Women in Science. In 2012, the Argentine Foundation of Nanotechnology (FAN) established a prize for Scientific Quality “Doctora Elizabeth Jares-Erijman,” and the newly established CONICET Institute of Nanosciences bears her name. 2.3 The Pervasive Influence of Gregorio Weber As has already been stated, Gregorio Weber plays a central role in this account, first of all due to the preeminent position he occupied in the scientific world. He is recognized as the person responsible for much of the theoretical and experimental developments in/of modern fluorescence spectroscopy. In particular, he pioneered the application of this technique in the biological sciences. In so doing and by virtue of his extraordinary human qualities, he served as an inspirational teacher to generations of spectroscopists and biophysicists working in basic science, biomedicine, and on industrial implementations of the numerous instrumentation and techniques developed in his lab. His list of achievements is extensive and unique: synthesis and application of small-molecule probes of hydrodynamic properties, polarity, and microviscosity; theory of fluorescence polarization and FRET; intrinsic fluorescence of the amino acids and of complexes of FAD and NADH; development of frequency domain fluorimetry; and studies of protein structure under pressure. In our work in G€ottingen, we were of course guided by the publications of this illustrious “father of biological fluorescence,” and a more personal association started in the 1970s. Gregorio and I shared an Argentine origin and thus it was not by accident that in 1993, during a visit to UIC, he introduced me to the Argentine husband–wife scientist pair Leo and Eli. Leo was Gregorio’s (last) postdoc and Gregorio recommended that both he and Eli extend their postdoctoral experience with a stint in our institute in G€ottingen, Germany. It was in this manner that Leo came to work with Bob Clegg and Eli came to work with me. In the three ensuing years, as well as thereafter, we kept “growing up with FRET” together, sharing our admiration of Gregorio Weber with other celebrated postdoctoral “FRETists” (now professors), such as Dorus Gadella Jr., Gerard Marriott, and Philippe Bastiaens. We also benefited from Gregorio’s insight and advice offered during occasional visits to the institute. Bob’s (and my) sentiments regarding Gregorio Weber are well expressed in Figure 2.1, taken from one of his many lectures on fluorescence and FRET, years later. Gregorio passed away, also of cancer, in 1997. 2.4 Contributions by Bob Clegg to FRET It is perhaps appropriate to invoke at this stage a particular formulation of the scientific method: (i) a scientific hypothesis can never be shown to be absolutely true, Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 12 Figure 2.1 Bob Clegg’s expression of indebtedness to Gregorio Weber. and must potentially be disprovable; (ii) the hypothesis must be useful until it is disproved; and (iii) the simplest hypothesis must be favored, unless it can be shown to be false (Ockham’s razor). Bob was meticulous about the application of these precepts and his contributions accordingly took two forms: direct (offering new insights) and indirect (providing inspiration to others). In his work he applied a keen innate physical intuition and was highly social in the sense of exhibiting an invariably pleasant, cheerful, and helpful disposition. Bob’s research interests can be summarized as follows: (i) development and applications of fluorescence lifetime-resolved imaging microscopy (FLI and FLIM) and the development of unique dedicated software for analysis of such data; (ii) development of instruments for rapid relaxation kinetics (T- and P-jump), and microsecond rapid mixing; and (iii) applications of the advanced instrumentation to an exceedingly wide range of biological systems, animal and vegetal. The underlying molecular mechanisms investigated included nucleic acid conformational equilibria and kinetics, and multisubunit functional proteins and photosynthetic systems. During his 21 years in Germany (1977–1998), Bob published 68 papers and contributions, most of which dealt with the structure of nucleic acids. Nine were devoted to the theory and practice of FRET and another nine were devoted to FLIM. Two FRET reviews had (and still have) a great impact on the field. The first [6] was a detailed blueprint for the FRET practitioner. The underlying theory was thoroughly presented as well as numerous techniques for the evaluation of population distributions and determinations of the FRET efficiency E: sensitized emission, donor quenching, decrease in donor lifetime, and changes in donor and acceptor anisotropy. In conclusion, Bob stated, “The measurements cannot be better than the molecules that are being measured, so extreme care j13 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 2.4 Contributions by Bob Clegg to FRET j 2 Remembering Robert Clegg and Elizabeth Jares-Erijman and Their Contributions to FRET Figure 2.2 FRET strategy (geometrical model) for determination of DNA helical parameters: (a) the D–A vector R “swings around” the helix as a function of separation N in the sequence and axial displacement L of the fluorophores relative to the bases to which they are attached, (b) FRET efficiency as a function of N (in base pair units). Inset: (ratio)A, a widely adopted FRET measure introduced by Bob; (ratio)A ¼ (acceptor emission sensitized by the donor and excited directly at the donor absorption band)/acceptor emission directly excited at its absorption band. Adapted from Ref. [9]. must be taken to ensure that the samples are well defined and pure. FRET with specifically labeled nucleic acids will surely become more popular in the near future; the method has the potential to distinguish many structural features and symmetries ranging over molecular distances less than 100 A.” How true! The second FRET review already dealt with FRET microscopy [7]. Bob contrasted in detail and unique clarity a quantum mechanical and two classical derivations of the FRET phenomenon, an issue he returned to repeatedly in later reviews and historical accounts (see later). A great deal of Bob’s research in Germany (and later, back in the United States) was devoted to the study of DNA helices, junctions, bulges, and kinks, much of it in close collaboration with David Lilley. FRET was an essential ingredient of this research, as is well illustrated in three publications that received wide attention [8–10]. One of these [9] featured the modulation of FRET according to the relative geometric disposition of FRET donor and acceptor positioned around a DNA helix (Figure 2.2). The development of FLI instrumentation was a key and long-range element of Bob’s research program, during both the German and subsequent US phases of his career. The emphasis was always on maximal speed and multiparametric acquisition, a pioneering example being the PhD thesis work of Peter Schneider [11]. Our extensive departmental involvement with FRET (flow cytometry and then imaging) and FLIM in the years leading up to the fall of the Berlin Wall was greatly facilitated by the participation of a large number of excellent colleagues from the Institute of Biophysics in Debrecen, Hungary. This circumstance was brought about by the farsighted efforts and perseverance of its director Sandor Damjanovich. Some of these individuals are featured in the group photograph taken at the symposium in Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 14 Figure 2.3 Bob Clegg (far left) and some former members of his group and other alumni of the Department of Molecular Biology, Max Planck Institute for Biophysical Chemistry, who attended the symposium honoring Theodor F€ orster in March 2011, G€ ottingen. honor of Theodor F€orster (“F€orster resonance energy transfer in life sciences”) in G€ottingen, March 2011, where some of us saw Bob for the last time (Figure 2.3). It was on this occasion that Bob managed to track down the house where F€ orster originated his theory and wrote his famous book Fluoreszenz organischer Verbindungen. Back in the United States (as of 1998) and established in an academic environment, Bob developed his capabilities to the full, including those as a gifted “gadgeteer.” He shared responsibility for the celebrated nationally funded research resource center, the LFD (Laboratory for Fluorescence Dynamics), established by Enrico Gratton in 1986 at UIC and relocated in 2006 to the University of California, Irvine. The LFD provided an excellent environment for cutting-edge technology development. Thus, in the decade of 2000–2010, a real-time field FLIM instrument, fully compatible with confocal optical configurations and with high contrast and sensitivity, was devised, allowing the high-speed acquisition of three-dimensional imaging and including spectral resolution. This instrumentation was applied to process prostate biopsies in an attempt to facilitate diagnosis of prostate cancer. In addition, the redistribution of a phototherapeutic/diagnostic compound (PpIX) in live tumor cells was investigated, just one activity establishing FLIM as an important technique in dermatology research [12]. A FRET redox biosensor was developed to measure the oxidation–reduction potentials in fluids and cells [13]. In parallel, a pressure-jump instrument was devised for studying photosynthetic plant systems, including living organisms such as algae [14]. The same method was applicable to kinetic studies of RNA/DNA conformational changes and binding of ligands. In fact, numerous post-2000 publications, most using FRET but not cited here, were dedicated to nucleic acid studies: four-way junctions (largely collaborations with Taekjip Ha and David Lilley), hammerhead ribozymes, protein–DNA interactions, j15 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 2.4 Contributions by Bob Clegg to FRET j 2 Remembering Robert Clegg and Elizabeth Jares-Erijman and Their Contributions to FRET ribosomal intersubunit dynamics, and probe–DNA interactions; the latter studies were conducted in the course of a very long-term collaboration with a close friend and colleague, Frank Loontiens. Data analysis was always a central issue for Bob in his work. As one example, the FLIM polar plot analysis of frequency domain FLIM image data was established [15] and extended by incorporating FRET-relevant spectral resolution [16] and newly created image analysis algorithms for selecting important image locations via their morphology using “wavelets” [17]. In addition, novel ways of “denoising” images depending on the type of noise (Poissonian or Gaussian) were devised [18], dramatically increasing the accuracy of the FLI. The potential of lifetime-resolved imaging in small organisms as well as live biological mammalian cells and photosynthesis (algae as well as higher plants) was to be enhanced for 3-D imaging using sample excitation single plane illumination microscopy (SPIM). Unfortunately, this work did not proceed beyond the planning stage due to Bob’s sickness. Bob did not neglect his dedication to the promotion of FRET history awareness and the applications of the technique, almost invariably coupled with FLIM. Accordingly, numerous reviews appeared (FRET [19–21] and FLIM [22–26]), which complemented and extended the earlier publications. Bob’s service to the scientific community was also evident in his many years as a member of the faculty of the long-standing Annual Workshop on FRET Microscopy, organized by Ammasi Periasamy, another prolific contributor to the FRET field and its literature. 2.5 Contributions by Eli Jares-Erijman to FRET Eli arrived in our lab in 1993 after a postdoc at UIC working in the lab of Ken Rinehart on the synthesis, isolation, and characterization of very complex natural products. She was an accomplished organic chemist but had had little exposure to fluorescence techniques, biophysical methods, and biomolecules. This situation changed in a very short time, such that 3 years later, Eli would return to Argentina as an accomplished biophysicist and an expert in fluorescence technology and probes. In fact, the latter served as primary objectives and motivators of this development, inasmuch as Eli displayed a keen ability to recognize the potential of new structural motifs, scaffolds, and mechanisms in creating innovative probes of molecular states, transitions, and localization. One of the first applications was in a study of noncanonical DNA such as Z-DNA and so-called parallel-stranded DNA (psDNA), using the FRET-based approach pioneered by Bob Clegg, while at the same time extending it conceptually and experimentally (new ratio functions). The left-handed character of Z-DNA was confirmed [27], but the helical sense of psDNA containing AA and GG base pairs and also presumed to be left-handed was not published because the extensive data posed (and still pose) problems of interpretation. Eli’s chemical acumen became very evident in the next FRET-based studies of photochromic compounds (diarylethenes) that provided a switchable acceptor function by virtue of dual (“open” and “closed”) states interconvertible by cycles Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 16 Figure 2.4 Photochromic FRET (pcFRET). Exposure of the diarylethene (open form) to near-UV light (250–320 nm) induces photocyclization to the closed form. The latter has an absorption band in the visible region overlapping with the emission band of a (the) donor, thus enabling FRET. Visible light leads to cycloreversion. The bistable diarylethenes exhibit little fatigue such that multiple cycles are feasible. of near-UV and visible light [28]. The distinctive absorption spectra of the closed and open forms differ in the degree of superposition with the emission of an appropriately selected donor, thereby leading to a change in the overlap integral J, one of the factors defining the F€orster transfer distance Ro. The mechanism of photochromic FRET (pcFRET) is depicted in Figure 2.4; it was explored by systematic structural modifications and careful thermodynamic and kinetic studies after Eli returned to Argentina [28–30]. At this juncture, Eli was operating as a Partner Group of our Max Planck Institute and was PI and co-PI on a number of nationally and internationally funded programs. Her major focus was on the development and application of smart sensors and devices combining luminescent, photochromic, and other small molecules with nanostructures such as quantum dots (QDs). In about 1998, QDs became commercially available through the auspices of Quantum Dot Corp. (QDC). Eli was probably the first person to conduct FRET experiments utilizing QDC QDs as donors and was instrumental in one of the first in-depth characterization of these new materials [31]. She readily perceived that emerging technologies based on a combination of chemistry, physics, and molecular biology were creating demand for smart materials serving as reporters and sensors in microand nanosystems [32]. The pioneering study of epidermal growth factor receptor (EGFR) activation and dynamics by microscopy of living cells using QD-EGF as ligands [33] was one of the first responses to this challenge, and stimulated numerous applications to other systems, including the insulin receptor [34,35]. Meanwhile, pcFRET was shown to operate at the level of a single particle [36] and to offer a new means for conducting isothermal relaxation kinetic measurements [37]. The pcFRET principle was extended to systems of core–shell QDs wrapped with an amphiphilic polymer containing photochromic groups and, in some j17 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 2.5 Contributions by Eli Jares-Erijman to FRET j 2 Remembering Robert Clegg and Elizabeth Jares-Erijman and Their Contributions to FRET constructs, a second fluorophore [38–40]. Such nanoparticles are water dispersible and can be reversibly modulated in fluorescence (quenched) by exercising the photochromic cycle. Other “J-engineered” QDs for sensing pH were based on FRET from the QD core to a shell of indicator molecules with pH-sensitive absorption spectra [41]. A quite different endeavor featured the use of “nanodevices” to sense and control the aggregation of amyloid proteins (specifically a-synuclein) in vitro and in the cellular context, thereby contributing to a better understanding of the molecular processes underlying the etiology of PD. This line of research led to the discovery of novel supramolecular intermediates in the aggregation pathway of a-synuclein preceding the formation of amyloid fibrils [42]. A prominent example is the “acuna” (amyloid “cradle”), a submicrometer structure that may be (at least partially) responsible for the toxicity and functional loss of dopaminergic neurons underlying PD [42]. The effort required a very interdisciplinary approach, combining numerous technologies such as organic synthesis, surface chemistry, physical and biophysical analysis, and quantitative microscopy to develop the sensors and apply them in context of cellular biology. Fluorogenic bisarsenical ligands [43], ratiometric [44] and/or solvatochromic probes [45], and NIR cyanines [46] also provided potential and actual novel FRET strategies for microscopy-based investigations of amyloid proteins in vitro and in living cells [47–51]. Current efforts in Eli’s as yet functional research group are also being directed at the design and synthesis of optimal photoswitchable probes for the emerging superresolution microscopies. As in the case of Bob Clegg, Eli published reviews on FRET imaging that have had a wide acceptance [52–54]. They are somewhat unusual in presenting novel views of photophysical phenomenon, such as the concept of a fluorophore as a photonic “enzyme” [52], and in offering an open-ended classification scheme for FRET methods. The latter include the donor and acceptor photobleaching techniques that originated from our FRET community in G€ ottingen, largely inspired by the unique publications, for example [55], of Tomas Hirschfeld, another illustrious member of the pantheon of spectroscopists. In publications [50,53,54], and in fact already in Ref. [27], it was proposed that in many FRET situations, calculations based on the ratio kt =kf may be preferable to the classical kt =kd , such that one can “bid farewell” to E and Ro . FRET is a moving target. 2.6 A Final Thought People survive in our memories if we keep them there by willfully recalling their personal qualities as well as their achievements. In Figure 2.5, we can appreciate that Bob and Eli, despite their distinctive ways and views, were two of a kind. We miss them both very much. I am greatly indebted to Bob Clegg’s family and other colleagues for material that made the writing of this chapter possible. Errors of commission and omission are mine alone. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 18 Figure 2.5 Eli and Bob on the grounds of the riverside campus of the University of Buenos Aires in 2010. References 1 F€ orster, T. (2012) Energy migration and 2 3 4 5 fluorescence. Journal of Biomedical Optics, 17, 011002. doi: 10.1117/1. JBO.17.1.011002 Weber, G. (1960) Fluorescence-polarization spectrum and electronic-energy transfer in proteins. The Biochemical Journal, 75, 345– 352. Steinberg, I.Z. (1971) Long-range nonradiative transfer of electronic excitation energy in proteins and polypeptides. Annual Review of Biochemistry, 40, 83–114. Stryer, L. (1978) Fluorescence energy transfer as a spectroscopic ruler. Annual Review of Biochemistry, 47, 819–846. Wu, P. and Brand, L. (1994) Resonance energy transfer: methods and applications. Analytical Biochemistry, 218, 1–13. 6 Clegg, R.M. (1992) Fluorescence resonance energy transfer and nucleic acids. Methods in Enzymology, 211, 353–388. 7 Clegg, R.M. (1996) Fluorescence resonance energy transfer (FRET), in Fluorescence Imaging Spectroscopy and Microscopy (eds X.F. Wang and B. Herman), John Wiley & Sons, New York, pp. 179–252. 8 Clegg, R.M., Murchie, A.I., Zechel, A., and Lilley, D.M. (1993) Observing the helical geometry of double-stranded DNA in solution by fluorescence resonance energy transfer. Proceedings of the National Academy of Sciences of the United States of America, 90, 2994–2998. 9 Gohlke F.C., Murchie, A.I., Lilley, D.M., and Clegg, R.M. (1994) Kinking of DNA and RNA helices by bulged nucleotides observed by fluorescence resonance energy transfer. Proceedings of the National j19 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 2 Remembering Robert Clegg and Elizabeth Jares-Erijman and Their Contributions to FRET 10 11 12 13 14 15 16 17 18 Academy of Sciences of the United States of America, 91, 11660–11664. Clegg, R.M., Murchie, A.I., and Lilley, D.M. (1994) The solution structure of the fourway DNA junction at low-salt conditions: a fluorescence resonance energy transfer analysis. Biophysical Journal, 66, 99–109. Schneider, P.C. and Clegg, R.M. (1997) Rapid acquisition, analysis, and display of fluorescence lifetime-resolved images for real-time applications. Review of Scientific Instruments, 68, 4107–4119. Hanson, K.M., Behne, M.J., Barry, N.P., Mauro, T.M., Gratton, E., and Clegg, R.M. (2002) Two-photon fluorescence lifetime imaging of the skin stratum corneum pH gradient. Biophysical Journal, 83, 1682–1690. Kolossov, V.L., Spring, B.Q., Clegg, R.M., Henry, J.J., Sokolowski, A., Kenis, P.J., and Gaskins, H.R. (2011) Development of a high-dynamic range, GFP-based FRET probe sensitive to oxidative microenvironments. Experimental Biology and Medicine (Maywood, NJ), 236, 681–691. Holub, O., Seufferheld, M.J., Gohlke, C., Govindjee, and Clegg, R.M. (2000) Fluorescence lifetime imaging (FLI) in real-time: a new technique in photosynthesis research. Photosynthetica, 38, 581–599. Redford, G.I. and Clegg, R.M. (2005) Polar plot representation for frequency-domain analysis of fluorescence lifetimes. Journal of Fluorine Chemistry, 15, 805–815. Chen, Y.C. and Clegg, R.M. (2011) Spectral resolution in conjunction with polar plots improves the accuracy and reliability of FLIM measurements and estimates of FRET efficiency. Journal of Microscopy, 244, 21–37. Buranachai, C., Kamiyama, D., Chiba, A., Williams, B.D., and Clegg, R.M. (2008) Rapid frequency-domain FLIM spinning disk confocal microscope: lifetime resolution, image improvement, and wavelet analysis. Journal of Fluorine Chemistry, 18, 929–942. Spring, B.Q. and Clegg, R.M. (2009) Image analysis for denoising full-field frequency-domain fluorescence lifetime images. Journal of Microscopy, 235, 221–237. 19 Clegg, R.M. (2005) Nuts and bolts of 20 21 22 23 24 25 26 27 28 29 excitation energy migration and energy transfer, in Chlorophyll a Fluorescence: A Signature of Photosynthesis (eds G.C. Papageorgiou and Govindjee), Springer, pp. 83–105. Clegg, R.M. (2006) The history of FRET: from conception through the labors of birth, in Reviews in Fluorescence 2006 (eds C.D. Geddes and J.R. Lakowicz), Springer, pp. 1–45. Clegg, R.M. (2009) F€orster resonance energy transfer - FRET what is it, why do it, and how it’s done, in FRET and FLIM Techniques: Laboratory Techniques in Biochemistry and Molecular Biology (ed. T.W. J. Gadella), Elsevier. Clegg, R.M., Holub, O., and Gohlke, C. (2003) Fluorescence lifetime-resolved imaging: measuring lifetimes in an image. Methods in Enzymology, 360, 509–542. Periasamy, A. and Clegg, R.M. (eds) (2009) FLIM Microscopy in Biology and Medicine, Taylor & Francis - Chapman & Hall/CRC Press. Noomnarm, U. and Clegg, R.M. (2009) Fluorescence lifetimes: fundamentals and interpretations. Photosynthesis Research, 101, 181–194. Chen, Y.C. and Clegg, R.M. (2009) Fluorescence lifetime-resolved imaging. Photosynthesis Research, 102, 143–155. Chen, Y.C., Spring, B.Q., and Clegg, R.M. (2012) Fluorescence lifetime imaging comes of age how to do it and how to interpret it. Methods in Molecular Biology, 875, 1–22. Jares-Erijman, E. and Jovin, T.M. (1996) Determination of DNA helical handedness by fluorescence resonance energy transfer. Journal of Molecular Biology, 257, 597–617. Giordano, L., Macareno, J., Song, L., Jovin, T.M., Irie, M., and Jares-Erijman, E.A. (2000) Fluorescence resonance energy transfer using spiropyran and diarylethene photochromic acceptors. Molecules, 5, 591–593. Jares-Erijman, E.A., Giordano, L., Spagnuolo, C., Kawior, J., Vermeij, R.J., and Jovin, T.M. (2004) Photochromic fluorescence resonance energy transfer (pcFRET): formalism, implementation, Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 20 30 31 32 33 34 35 36 and perspectives. Proceedings of SPIE, 5323, 13–26. Giordano, L., Jovin, T.M., Irie, M., and Jares-Erijman, E.A. (2002) Diheteroarylethenes as thermally stable photoswitchable acceptors in photochromic fluorescence resonance energy transfer (pcFRET). Journal of the American Chemical Society, 124, 7481–7489. Grecco, H.E., Lidke, K.A., Heintzmann, R., Lidke, D.S., Spagnuolo, C., Martinez, O.E., Jares-Erijman, E.A., and Jovin, T.M. (2004) Ensemble and single particle photophysical properties (two-photon excitation, anisotropy, FRET, lifetime, spectral conversion) of commercial quantum dots in solution and in live cells. Microscopy Research and Technique, 65, 169–179. Jares-Erijman, E.A., Spagnuolo, C., Giordano, L., Etchehon, M., Kawior, J., Ma~ nalich-Arana, M., Bossi, M., Lidke, D.S., Post, J.N., Vermeij, R.J., Heintzmann, R., Lidke, K.A., Arndt-Jovin, D.J., and Jovin, T. M. (2004) Novel (bio)chemical and (photo) physical probes for imaging live cells, in Supramolecular Structure and Function, vol. 8 (ed. G. Pifat-Mrzljak), Kluwer, Amsterdam, pp. 99–118. Lidke, F.D.S., Nagy, P., Heintzmann, R., Arndt-Jovin, D.J., Post, J.N., Grecco, H., Jares-Erijman, E.A., and Jovin, T.M. (2004) Quantum dot ligands provide new insights into erbB/HER receptor-mediated signal transduction. Nature Biotechnology, 22, 198–203. Giudice, J., Jares-Erijman, E.A., and Leskow, F.C. (2013) Endocytosis and intracellular dissociation rates of human insulin–insulin receptor complexes by quantum dots in living cells. Bioconjugate Chemistry, 24, 431–442. Giudice, J., Barcos, L.S., Guaimas, F.F., Penas-Steinhardt, A., Giordano, L., JaresErijman, E.A., and Coluccio Leskow, F. (2013) Insulin and insulin like growth factor II endocytosis and signaling via insulin receptor B. Cell Communication and Signaling, 11, 18. Jares-Erijman, E.A., Giordano, L., Spagnuolo, C., Lidke, K.A., and Jovin, T.M. (2005) Imaging quantum dots switched on and off by photochromic Fluorescence 37 38 39 40 41 42 43 44 resonance energy transfer (pcFRET). Molecular Crystals and Liquid Crystals, 430, 257–265. Jovin, T.M. and Jares-Erijman, E.A. (2005) Photochromic relaxation kinetics (pcRelKin). Molecular Crystals and Liquid Crystals, 430, 281–286. Díaz, S., Menendez, G., Etchehon, M., Giordano, L., Jovin, T.M., and JaresErijman, E.A. (2011) Photoswitchable water-soluble quantum dots: pcFRET based on amphiphilic photochromic polymer coating. ACS Nano, 5, 2795–2805. Díaz, S., Giordano, L., Jovin, T.M., and Jares-Erijman, E.A. (2012) Modulation of a photoswitchable dual-color quantum dot containing a photochromic FRET acceptor and an internal standard. Nano Letters, 12, 3537–3544. Díaz, S.A., Giordano, L., Azcarate, J.C., Jovin, T.M., and Jares-Erijman, E.A. (2013) Quantum dots as templates for selfassembly of photoswitchable polymers: small, dual-color nanoparticles capable of facile photomodulation. Journal of the American Chemical Society, 135, 3208–3217. Menendez, G., Roberti, M.J., Sigot, V., Etchehon, M., Jovin, T.M., and JaresErijman, E.A. (2009) Interplay of multivalency and optical properties of quantum dots: implications for sensing and actuation in living cells. Proceedings of SPIE, 7189, 71890-P1–71890-P9. Fauerbach, J.A., Yushchenko, D.A., Shahmoradian, S.H., Chiu, W., Jovin, T.M., and Jares-Erijman, E.A. (2012) Supramolecular non-amyloid intermediates in the early stages of a-synuclein aggregation. Biophysical Journal, 102, 1127–1136. Spagnuolo, C.C., Massad, W., Miskoski, S., Menendez, G.O., Garcia, N.A., and JaresErijman, E.A. (2009) Photostability and spectral properties of fluorinated fluoresceins and their biarsenical derivatives: a combined experimental and theoretical study. Photochemistry and Photobiology, 85, 1082–1088. Yushchenko, D.A., Fauerbach, J.A., Thirunavukkuarasu, S., Jares-Erijman, E. A., and Jovin, T.M. (2010) Fluorescent ratiometric MFC probe sensitive to the early stages of a-synuclein aggregation. j21 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 2 Remembering Robert Clegg and Elizabeth Jares-Erijman and Their Contributions to FRET 45 46 47 48 49 Journal of the American Chemical Society, 132, 7860–7861. Giordano, L., Shvadchak, V.V., Fauerbach, J.A., Jares-Erijman, E.A., and Jovin, T.M. (2012) Highly solvatochromic 7-arylhydroxychromones. The Journal of Physical Chemistry Letters, 3, 1011–1016. Menendez, G.O., Pichel, M.E., Spagnuolo, C.C., and Jares-Erijman, E.A. (2013) NIR fluorescent biotinylated cyanine dye: optical properties and combination with quantum dots as a potential sensing device. Photochemical & Photobiological Sciences, 12, 236–240. Roberti, M.J., Bertoncini, C.W., Klement, R., Jares-Erijman, E.A., and Jovin, T.M. (2007) Fluorescence imaging of amyloid formation in living cells by a functional, tetracysteine-tagged a-synuclein. Nature Methods, 4, 345–351. Roberti, M., Morgan, M., Menendez, G., Pietrasanta, L., Jovin, T.M., and JaresErijman, E.A. (2009) Quantum dots as ultrasensitive nanoactuators and sensors of amyloid aggregation in live cells. Journal of the American Chemical Society, 131, 8102–8107. Roberti, M.J., Jovin, T.M., and JaresErijman, E.A. (2011) Confocal fluorescence 50 51 52 53 54 55 anisotropy and FRAP imaging of a-synuclein amyloid aggregates in living cells. PloS One, 6, e23338. Roberti, M.J., Giordano, L., Jovin, T.M., and Jares-Erijman, E.A. (2011) FRET imaging by kt/kf. ChemPhysChem, 12, 563–566. Roberti, M.J., F€olling, J., Celej, M.S., Bossi, M., Jovin, T.M., and Jares-Erijman, E.A. (2012) Imaging nanometer-sized a-synuclein aggregates by superresolution fluorescence localization microscopy. Biophysical Journal, 102, 1598–1607. Jares-Erijman, E.A. and Jovin, T.M. (2003) FRET imaging. Nature Biotechnology, 21, 1387–1395. Jares-Erijman E.A. and Jovin, T.M. (2006) Imaging molecular interactions in living cells by FRET microscopy. Current Opinion in Chemical Biology, 10, 1–8. Jares-Erijman, E.A. and Jovin, T.M. (2009) Reflections on FRET imaging: formalism, probes, and implementation, in FRET and FLIM Imaging Techniques (ed. T. Gadella Jr.), Academic Press, pp. 475–517. Hirschfeld, T. (1976) Quantum efficiency independence of the time integrated emission from a fluorescent molecule. Applied Optics, 15, 3135–3139. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 22 3 F€ orster Theory B. Wieb van der Meer 3.1 Introduction Theory without experiment is like the sound of one hand clapping. F€ orster theory is not like that at all. It is a thundering ovation linking theory and experiment by explaining the relationship between spectral overlap, energy transfer, and proximity. This chapter explains F€orster’s contributions to the theory of resonance energy transfer. The readers of this chapter form, no doubt, a highly diverse group of people. Most readers are probably only interested in the bottom line. Others may want to know details. But which details? There are so many. To help students and specialists find what they need, the chapter is presented as a sequence of a large number of sections that are short and focused. 3.2 Pre-F€ orster This section is based on some of the information in the most popular papers by F€orster [1–5], Chapter 5 of Ref. [6], and Clegg’s history of FRET [7]. The emphasis here is on the contributions of F€orster’s predecessors and contemporaries. If you want to know who the scientists were who inspired F€ orster and what the science was that motivated him, you should read his most important papers. His most important, that is, his most cited papers are his papers published in 1946 [1] 1948 [2], and 1949 [4] and reviews published in 1959 [3] and 1965 [5]. F€ orster’s papers are not easy to understand. The language is not a problem because four of the five are in English or translated into English. They are difficult because they use a lot of math and complicated spectroscopic concepts. Nevertheless, if you are serious about FRET, you should study them. Start with his 1946 paper [1] and the 1959 review [3]. These papers are much more readable than F€orster’s most cited paper [2], because his 1946 paper presents a very clear verbal description of the essential ideas on which the theory is based and a thorough review of the experimental evidence of the importance of resonance and his 1959 paper is designed to provide a conceptual understanding of the FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j23 j 3 F€orster Theory FRET phenomenon with a minimum of equations, whereas his 1948 paper is a rigorous treatment of the theory. In Ref. [3], he gives an overview of the then available literature. For example, he mentions experiments proving that the observed transfer mechanism could not be due to trivial reabsorption. His discussion of the various transfer mechanism is very interesting. He compares, in Table 1 in Ref. [3], FRET, reabsorption, donor–acceptor complex formation, and collisional quenching. This table has been adopted in a slightly modified form in Section 3.7. In Refs [1,2], he focuses on the theory of homotransfer, FRET between like molecules, whereas in Ref. [4] (only available in German), he emphasizes heterotransfer, FRET between unlike molecules. If one is interested in the fundamental aspects of the theory, refer to Refs [1,2,5]. Reference [5] is a review. In this paper, he does more than just rehash his FRET theory. Of the 10 sections, only the last one is about FRET. In his 1965 paper, he also discusses exciton theory, strong coupling, weak coupling, and very weak coupling (very weak coupling is the basis for FRET). His 1965 paper introduces an extension of his theory put forward in his 1948 paper. This extension allows a description of the time dependence of the donor fluorescence and the relation between the quantum yield of the donor and the acceptor concentration (see Sections 3.16 and 3.17). Newton said, “If I have seen further it is only by standing on the shoulders of giants” [8]. Who were the giants for F€orster? J. Perrin and F. Perrin, Cario, Franck, Kallmann, and London. In the late 1940s when F€ orster started his work on energy transfer, the phenomenon of sensitized fluorescence was well established [2–5,7]. Cario had shown in 1922 that transfer of energy had taken place from excited mercury atoms to thallium atoms in a mercury–thallium vapor mixture [9]. Cario and Franck had presented similar results in a mercury–hydrogen system [10]. Many other experimentalists had presented evidence for sensitized fluorescence from the vapors of silver, cadmium, lead, zinc, and indium in the presence of mercury vapor [7,9,10]. The starting point for a theoretical framework showing the role of resonance in energy transfer was Franck’s principle: If effective energy transfer is to take place from initially excited molecules to quenching molecules, the excited states of the quenchers must be in energy resonance with the primarily excited states [11]. Kallmann and London [7,12] proposed a theory of energy transfer that can be considered to be a precursor of F€orster’s theory in that it is based on the idea of resonance and has the correct distance dependence of the transfer efficiency. However, unrealistically sharp spectra were assumed and the link between the distance dependence and spectra had limited significance. J. Perrin [13,14] introduced a classical theory modeling the fluorophores as electrical dipoles oscillating at a single frequency with a rate of transfer proportional to the inverse of the distance to the third power, not the sixth power as F€orster later found. As a result, the predicted distance over which energy transfer would take place is much too large [7]. F. Perrin [15,16] designed a quantum mechanical version of this theory extending the work by Kallmann and London to transfer in solutions. The molecules are assumed to have two states, a ground state and an excited state, so that the spectra would show sharp peaks. F. Perrin found the same distance dependence as J. Perrin did. However, F. Perrin did invoke collision broadening of the spectra by the solvent, decreasing the predicted range of transfer, but it was still too large [7]. An interesting overview of the Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 24 contributions of J. Perrin and F. Perrin, father and son, is available [17]. Oppenheimer and Arnold [7,18,19] pointed out that the phenomenon of resonance energy transfer is very similar to internal conversion in radioactivity where an excited nucleus transfers energy without radiation to one of the orbital electrons resulting in ejection of this electron. Using this similarity, they derived an expression of the rate of energy transfer for the case where the acceptors are randomly distributed around a donor. Clegg showed that modifying their expression for the rate of transfer from one dominant frequency to a spectrum of frequencies leads to F€ orster’s famous equation for the donor–acceptor distance at which the rate of transfer equals that of donor emission [7] (see the equation forR0 in Section 3.3). However, the fact is that Oppenheimer and Arnold did not make these modifications. They did not come up with the idea to incorporate experimentally obtained spectra into their theory. F€orster did [1–5]. This is what sets F€orster apart from his predecessors and contemporaries. They all assumed one dominant frequency and ignored experimental data on spectra. F€orster’s most important innovation was to incorporate experimentally obtained parameters such as spectra, quantum yield, and lifetimes into his theory, making it refutable, accessible, and extremely useful. 3.3 Bottom Line To observe FRET, the following conditions must be met: 1) Donor and acceptor must have strong electronic transitions in the UV, visible, or IR. 2) Spectral overlap must exist between donor emission and acceptor absorbance (see Section 3.5). 3) Donor and acceptor must be close, but not too close (see Sections 3.6 and 3.7). 4) The orientation factor should not be too small (see Section 3.8). 5) The donor emission should have a reasonably high quantum yield (see Equation 3.3 and also data in Chapter 14). The following are the key quantities in F€orster’s theory: kT ¼ rate of energy transfer (see Equation 3.2). tD ¼ lifetime of the donor excited state in the absence of acceptor (see Equations 3.1 and 3.2). r DA ¼ distance between donor and acceptor. R0 ¼ F€orster distance, that is, the donor–acceptor distance at which kT ¼ 1=tD , so that at that particular distance, the probability of the excited donor to fluoresce is equal to the probability of transfer of energy to the acceptor (see Equations 3.3 and 3.3a–3.3c). E ¼ efficiency of transfer (see Equation 3.1). J ¼ overlap integral (see Section 3.5 and Equation 3.4). k2 ¼ orientation factor (see Section 3.8 and Equation 3.3). WD ¼ quantum yield of the donor fluorescence in the absence of acceptor (see Equation 3.3). n ¼ refractive index of the medium (see Equation 3.3). j25 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.3 Bottom Line j 3 F€orster Theory Constants in F€orster’s theory are as follows: p ¼ 3:141592654; ln 10 ¼ 2:302585093; orster used N 0 N A ¼ 6:0221415 1023 per mole ðactually; F€ ¼ 6:0221415 1020 per millimoleÞ: The conclusion of F€orster’s theory can be conveniently written as the following set of three equations: kT R6 ¼ 6 06 : kT þ 1=tD R0 þ r DA 1 R60 : kT ¼ tD r 6DA E¼ R60 ¼ 9ðln 10Þk2 WD J : 128p5 n4 N A ð3:1Þ ð3:2Þ ð3:3Þ For a derivation of these equations from classical theory, see Sections 3.5–3.14 (Sections 3.5–3.7 introduce basic ideas, the derivation starts in Section 3.8), and from quantum mechanical theory, see Section 3.15. Note that the second equation follows from the first (and the first from the second). An alternative expression for the rate of transfer kT is 2 k 1 CDA ; kT ¼ 6 ð3:2aÞ n4 r DA with CDA ¼ 9ðln 10ÞWD J : 128p5 tD N A ð3:2bÞ This formulation, which is often used in photosynthesis, establishes a clear separation between spectral properties (CDA ), geometric properties (k2 =r 6DA ), and environmental factors (n) [20]. When r DA is in nanometers and kT is in inverse picoseconds, C DA is expressed in nm6/ps. Figure 3.1 illustrates the relations between efficiency and donor–acceptor distance and F€orster distance and between F€orster distance and kappa-squared, overlap integral, refractive index, and quantum yield. 3.4 9000-Form, 9-Form, and Practical Expressions of the R60 Equation F€orster used N 0 instead of N A in Ref. [2], but used N A with 9000 instead of 9 in Ref. [3]. However, N 0 ¼ N A as both have a unit and represent the same amount of particles per mole: N 0 ¼ 6:02 1020 mmol1 ¼ 6:02 1023 mol1 ¼ N A . Braslavsky et al. pointed out that the frequently quoted 9000-form of Equation 3.3 (with a factor of 9000 instead of 9) is incorrect [21]. Simplifying Equation 3.3 by substituting Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 26 Figure 3.1 (a) The transfer efficiency E versus the donor–acceptor distance r DA between 0 and 2 times the F€ orster distance. (b) The transfer efficiency E versus the F€ orster distance R0 between 0 and 2 times the donor–acceptor distance. (c) The relative F€ orster distance versus the orientation factor, k2 , which varies orster distance between 0 and 4. R0 [2/3] is the F€ orster distance at k2 ¼ 2=3. (d) The relative F€ R0/R0(J/J1) versus the relative overlap integral J/J1, where J is the overlap integral and J1 is a typical value of this integral, say 1 OLI. orster distance that at J ¼ J1. R0( J ¼ J1) is the F€ J-values vary over a wide range (see Chapter 14). (e) The relative F€ orster distance versus n, the refractive index of the medium in which the donor and acceptor are embedded. R0(n ¼ 1.4) is the F€ orster distance for the refractive index equal to 1.4. All refractive index values in the literature are in the 1.33–1.6 range. The values 1.34 and 1.6 are the ones used most frequently. (f) The relative F€ orster distance versus FD , the quantum yield of the donor in the absence of orster the acceptor R0 (FD ¼ 0:5) is the F€ distance at FD ¼ 0:5. FD varies between 0 and 1 (see Chapter 14 for data). j27 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.4 9000-Form, 9-Form, and Practical Expressions of the R06 Equation j 3 F€orster Theory in the constants with units shown explicitly yields [21] 2 1=6 R0 k WD Jl : ¼ 0:02108 nm n4 mol1 dm3 cm1 nm4 ð3:3aÞ Equation 15 in Ref. [22] and the expressions given on page 16 of Ref. [6] are equally valid, but the expression (3.3a) is preferable as it has the advantage that the choice of units is absolutely clear, so that mistakes can be easily avoided. In practice, the OLI (overlap–integral unit, introduced in Chapter 7 of Ref. [6]) is convenient. This unit is defined as OLI ¼ 1014 mol1 dm3 cm1 nm4 ¼ 1014 mol1 dm3 cm3 , and is used in Chapter 14. Alternative forms of (3.3a) using the OLI are as follows: 2 1=6 R0 k WD 100J l : ¼ 2:108 nm n4 OLI ð3:3bÞ 2 1=6 R0 k WD J l ¼ 4:542 : nm n4 OLI ð3:3cÞ The following is a step-by-step derivation of Equation 3.3a from Equation 3.3: 1) Divide both sides of Equation 3.3 by nm6 and substitute in all the constants, including the unit of Avogadro’s number, and Equation 3.3 thus becomes ( ) R60 9ð2:302585Þ k2 WD J : ¼ nm6 n4 nm6 128ð306:0197Þð6:022 1023 Þmol1 1 2) Move to J and convert nm6 to dm3 cm1 nm4 using 17 mol over 2 3 10 nm = dm cm1 ¼ 1. 2 17 k WD J 10 nm2 : n4 dm3 cm1 mol1 nm6 2 R6 k WD J 3) Simplify yielding: 0 6 ¼ 87:8533 1012 : nm n4 mol1 dm3 cm1 nm4 R60 ¼ nm6 9ð2:302585Þ 128ð306:0197Þð6:022 1023 Þ 4) Take the sixth root and arrive at Equation 3.3a. 3.5 Overlap Integral The overlap integral can be calculated using wavelength (3.4), wave number (3.6), or frequency (3.8). Most frequently, the wavelength form is used. This is often referred to as J l and defined as: ð J l ¼ J ¼ f D ðlÞeA ðlÞl4 dl: ð3:4Þ The integral in Equation 3.4 extends over the region that encompasses the line shapes of the relevant donor emission and acceptor absorption bands. Extending the Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 28 Figure 3.2 (a) Donor emission spectrum (blue curve) and acceptor extinction spectrum (red curve) versus wavelength. The green area indicates that there is overlap between the two spectra. The green area is not the overlap integral. (b) This schematic graph has three different vertical scales, one in nm1 for the blue curve (normalized donor fluorescence, the area under this curve is one), another in M1 cm1 for the red curve (acceptor extinction), and the third scale is in OLI/nanometer for the green curve (overlap curve). The area under the green curve is the overlap integral. integrals from zero to infinity may add irrelevant addenda to the integrals when other areas of overlap occur at very high and/or low wavelengths not relevant to the specific energy transfer process of interest.1) The overlap integral J ¼ J l is conveniently expressed in OLIs, l is the wavelength of the light, most often expressed in nanometers, eA ðlÞ is the molar extinction coefficient of the acceptor, usually in M1 cm1, and f D ðlÞ is the fluorescence spectrum of the donor normalized on the wavelength scale: f D ðlÞ ¼ Ð F Dl ðlÞ ; F Dl ðlÞdl ð3:5Þ where F Dl ðlÞ is the donor fluorescence per unit of wavelength interval and the integral extends over the relevant donor emission band(s). A schematic illustration of overlap is Figure 3.2. The green area in Figure 3.2a is not the overlap integral. The green area only shows that there is overlap and is certainly not a reliable measure for the magnitude of the overlap integral. The donor emission f D ðlÞ (blue curve in Figure 3.2) is usually expressed in 1/nm. The acceptor extinction eA ðlÞ (red curve) is in M1 cm1, and the overlap curve (green) is in OLI/nm. The area under the green curve in Figure 3.2b is the overlap integral. There are three different vertical scales in Figure 3.2b. Therefore, by adjusting one scale with respect to the others, the appearance of this figure can be adjusted. However, whatever scale adjustment is made, Figure 3.2a can never be made to resemble Figure 3.2b, because the 1) Andrews, D.L. (2012) Private communication. j29 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.5 Overlap Integral j 3 F€orster Theory wavelength to the fourth factor will cause the peak of the overlap curve to be always at a wavelength larger than the one where donor emission and acceptor extinction intersect. The overlap integral in wave number form is ð f ð~nÞeA ð~nÞ J¼ D 4 d~n: ð3:6Þ ~n Here, the integral extends over the region that encompasses the relevant donor emission and acceptor absorption bands in terms of wave number, ~ n ¼ 1=l, conveniently expressed in 1/nm, eA ð~nÞ is the molar extinction coefficient of the acceptor, usually in M1 cm1, and f D ð~nÞ is the fluorescence spectrum of the donor normalized on the wave number scale: F D~n ð~nÞ ð~nÞd~n; f D ð~nÞ ¼ Ð F D~n ð3:7Þ where F D~n ð~nÞ is the donor fluorescence per unit of wave number interval and the integral extends over the relevant donor emission band(s). The overlap integral in frequency form is the one appearing in F€ orster theory (see Section 3.12): ð f ðnÞeA ðnÞ J ¼ c4 D 4 dn: ð3:8Þ n Here, the integral extends over the region that encompasses the relevant donor emission and acceptor absorption bands in terms of frequency, n ¼ c=l, c is the speed of light in vacuo, which is equal to about 3 108 m=s (more precisely c ¼ 2:99792458 108 m=s). The frequency is conveniently expressed in hertz (¼1/s), eA ðnÞ is the molar extinction coefficient of the acceptor, usually in M1 cm1, and f D ðnÞ is the fluorescence spectrum of the donor normalized on the frequency scale: f D ðnÞ ¼ Ð F Dn ðnÞ ; F Dn ðnÞdn ð3:9Þ where F Dn ðnÞ is the donor fluorescence per unit of frequency interval, and the integral extends over the relevant donor emission band(s). At first sight, the conversions from Equations 3.4–3.6, and to 3.8 look inconsistent with the rules of calculus, but they are actually correct and follow from the “first law of photophysics” (Chapter 2 of Ref. [6]): F Dl dl ¼ F D~n d~n ¼ F Dn dn: ð3:10Þ The intensities are measured using monochromators or filters with a certain resolution or bandwidth. The reading on the instrument is proportional to this bandwidth if it is sufficiently small. The intensity is taken to be proportional to the reading per wavelength or wave number or frequency. This proportionality is the idea behind the first law of photophysics. Note that the extinction coefficient transforms as eA ðlÞ ¼ eA ð~nÞ ¼ eA ðnÞ, because it is proportional to the logarithm of a ratio of intensities. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 30 The overlap integral depends on both the absorption spectrum of the acceptor and the emission spectrum of the donor. Absorption spectra vary as a rule relatively little with change of solvent or temperature, but emission spectra may be very sensitive to environment, a well-studied example being the fluorescence of indole derivatives [23]. Analytical results for the overlap integral have been derived for bands of Gaussian and log normal line shapes [24]. 3.6 Zones F€orster visualized a donor as a group of electrical oscillators close together. These electrical oscillators produce an electrical field in the space around the donor. This space consists of four zones: the contact zone or Dexter zone [25], the near zone or the near field, the intermediate zone, and the far zone (also called the far-field or the radiation zone). The concept of zones, illustrated in Figure 3.3, dates back to Hertz [7] who actually considered three zones: the near, intermediate, and far, because he set out to confirm Maxwell’s prediction of electromagnetic waves [7] and was not interested in distances very close to the electrical oscillators. The zones can be defined in terms of a distance b: b¼ l ; 2pn Figure 3.3 The space around a donor fluorophore can be divided into four zones. These zones are shown here on a logarithmic scale with the outer radius of each ring being a factor of 10 larger than the inner radius. The donor occupies the center of the contact zone, which extends up to about a nanometer or more depending on the donor size (see Table 3.1). Around this zone is the near field, about 1–10 nm from the donor. The near field is the only zone where F€ orster theory applies. ð3:11Þ Around the near field is the intermediate zone from 10–1000 nm. Outside the intermediate zone is the far field where electromagnetic radiation takes place. If the acceptor concentration is sufficiently small – so that the probability of finding an acceptor in the contact zone is very small – and the sample is not too large – so that the probability of reabsorption is small, FRET is the dominant mode of energy transfer. j31 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.6 Zones j 3 F€orster Theory where l is the wavelength of the donor fluorescence that is usually in the 300– 800 nm range and n is the index of refraction of the medium in which the donor and acceptor are embedded [26]. This index has values typically between 1.3 and 1.6. So, b is about 100 nm or a little smaller. Properties of these zones are listed in Table 3.1. FRET happens in the near field, that is, roughly in the 1–10 nm range. If it is less than about 1 nm, F€orster theory does not apply for at least two reasons: First, the complex formation may occur between donor and acceptor at such a proximity (see also Section 3.7 and Refs [3,20,25]). Second, the F€ orster’s theory is based on the ideal dipole approximation (IDA) and the IDA breaks down if the donor–acceptor distance is on the order of 1 nm [28]. If the distance is larger than about 10 nm, contributions to the electric field that are ignored in F€ orster’s theory become relevant. In the radiation zone, the acceptor is capable of reabsorbing light emitted by the donor (see Section 3.7). Table 3.1 Zones around the donor. Name Alternative name Inner radius Outer radius Characteristics Contact zone Dexter zone 0 0.01b (1 nm) Near-field zone Near field 0.01b (1 nm) 0.1b (10 nm) 0.1b (10 nm) 10b (1000 nm) 10b Infinite The ideal dipole approximation breaks down [25]. An acceptor in this zone may form a complex with the donor [3]. F€orster theory does not apply. For larger molecules (chlorophyll and porphorin), the outer radius is about 2–3 nm [27]. The distance dependence of transfer is reviewed in Ref. [20] F€orster theory is valid only in this zone. The electric field due to oscillating donor charges can be considered as a sum of dipole terms with a 1/distance3 dependence. The inner radius may be bigger for larger molecules (see above) The electric field due to oscillating donor charges has three terms with different distance dependence, none of which is dominant. F€orster theory does not apply Electromagnetic donor emission takes place in this zone. The electric field due to oscillating donor charges has a 1/distance dependence. The electric field lines are pinched off and transverse waves are formed [7]. F€orster theory does not apply. Reabsorption will occur Intermediate zone Radiation zone Far field Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 32 3.7 Transfer Mechanisms F€orster was well aware of the fact that there are at least four different mechanisms by which excitation energy can be transferred from a donor to an acceptor. These are resonance energy transfer, reabsorption, complex formation, and collision quenching. Resonance energy transfer (also called FRET2)) [29], the main topic of this book, is the radiationless transmission of an energy quantum from its site of absorption to the site of its utilization in a molecule, or system of molecules, by resonance interaction between chromophores, over distances considerably greater than interatomic without conversion to thermal energy and without the donor and acceptor coming into kinetic collision. Reabsorption or “trivial reabsorption” is the emission of a photon by the donor with the subsequent absorption of that photon by the acceptor. Complex formation is the creation of an excited-state complex of a donor and an acceptor that are in proximity, essentially in molecular contact with each other. Collision quenching can occur when an excited molecule loses its excitation energy to another molecule by colliding with it. As F€ orster pointed out, these four transfer mechanisms have different characteristics and can, therefore, be distinguished experimentally [3]. Table 3.2, adopted with minor modifications, from one of F€orster’s papers [3], summarizes these different characteristics. Quantum Table 3.2 Characteristics of transfer mechanisms. Sample volume: with increasing volume, transfer exhibits Viscosity: with increasing viscosity, transfer exhibits Donor lifetime: because of transfer, the donor lifetime shows Donor fluorescence spectrum: comparing transfer and no transfer. This spectrum is Donor absorption spectrum: comparing transfer and no transfer. This spectrum is Resonance energy transfer Reabsorption Complex formation Collision quenching No change Increase No change No change No change No change No change Decrease Decrease No change No change Decrease Unchanged Changeda) Unchanged Unchanged Unchanged Unchanged Changeda) Unchanged a) Changes only apply to the wavelength, not to the intensity. 2) There is general agreement about the FRET acronym. However, there is no consensus yet about the meaning of the letters in FRET. Many authors read it as “fluorescence resonance energy transfer,” while many others as “F€orster resonance energy transfer.” The author of this chapter prefers “fluorescence with resonance energy transfer.” “Fluorescence-detected resonance energy transfer” was proposed by Vanbeek et al. [29]. j33 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.7 Transfer Mechanisms j 3 F€orster Theory electrodynamics teaches that all electric and magnetic interactions are mediated by photons even in the near field. The photons in the near field are actually virtual photons. As a result, FRET and trivial reabsorption can be interpreted as two different aspects of the same phenomenon (see Section 3.19) [30]. 3.8 Kappa-Squared Basics The previous sections can be considered to provide an introduction to F€ orster theory. This is the first of a group of sections (Section 3.8–3.14) forming F€ orster’s classical derivation of his equations (Equations 3.1–3.3). The orientation factor, kappa-squared, is the square of k, which is defined as k ¼ cos qT 3cos qD cos qA : ð3:12Þ Here, qD is the angle between the donor emission transition moment and the donor–acceptor connection line, qA is the angle between the acceptor absorption transition moment and the donor–acceptor connection line, and qT is the angle between the donor emission transition moment and the acceptor absorption transition moment. Kappa-squared varies between 0 and 4 and is discussed, in more detail, in Chapter 4. The relation between the F€ orster distance and the kappasquared is shown in Figure 3.1c. In F€orster theory, k appears in terms of dot products ^ ^a, and ^r , which are unit vectors: d^ along the donor dipole, ^ between d, a along the acceptor dipole, and ^r along the line from the donor to the acceptor. ^ d ^r ¼ cos qD , ^a ^r ¼ cos qA , and d^ ^a ¼ cos qT . Kappa in terms of dot products is k ¼ ^d ^a 3 ^d ^r ð^r ^aÞ: ð3:13Þ The angles and unit vectors are illustrated in Figure 3.4. The amplitude of the donor dipole moment is qe DD (qe is the charge of an electron d and DD is the displacement of the charge, both shown below). It oscillates along the ^ direction at frequency nDONOR (in general, there is a distribution of frequencies) and ^ ^a, and ^r . Figure 3.4 Illustration of the angles uD , uA , and uT and the unit vectors d, Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 34 generates an electric field at the location of the acceptor. Donor and acceptor are embedded in a medium with refractive index n, with a distance r DA between them. At time t, the electric field generated by the donor dipole at the location of the acceptor, in the near field, is given by h i qe DD 3 ^d ^r ^r ^d cos ð2pnDONOR tÞ ~ E ¼ : ð3:14Þ 4pe0 n2 r 3DA The component of this field along the acceptor dipole is q DD kcos ð2pnDONOR tÞ ~ E ^a ¼ e : 4pe0 n2 r 3DA ð3:15Þ The mks unit system is used here. In the cgs system, the 4pe0 factor is replaced by 1. 3.9 Ideal Dipole Approximation The donor is a group of oscillating charges. We can imagine a sphere drawn around the center of the donor with radius RD containing all these charges. Similarly, the acceptor is a group of oscillating charges contained in a sphere around its center with radius RA . In the near field, the distance r DA between the center of the donor and that of the acceptor is much larger than RD and also much larger than RA . As a result, the ideal dipole approximation holds: The electromagnetic interaction between donor and acceptor is a dipole–dipole interaction, and all interactions due to higher multipoles can be ignored [31]. The ideal dipole approximation is illustrated in Figure 3.5. It must be emphasized that the relevant donor and acceptor dipoles are not permanent dipoles but oscillating dipoles; in quantum mechanical terms, they are transition dipoles. Consistent with this dominance of the dipole moment above all other multipoles, F€orster visualized a donor or an acceptor molecule as a group of coupled electrical oscillators [32]. Each electrical oscillator consists of an electron elastically bound to a nucleus. The nucleus is stationary,3) but the electron can oscillate along a certain direction (not along other directions). The charge of the electron is qe (qe ¼ 1:60217646 1019 C) and its mass is me (me ¼ 9:10938188 1031 kg). The values qe and me do not appear in the final results because they are taken up by the spectral properties of the donor and the acceptor. The donor dipole is ^ a unit vector, and dipole situated at the center of the donor and has a direction d, moment~ p D ¼ qe DD ^d, where DD ^d is a vector sum of fluctuating vectors oscillating at a range of frequencies. The vector DD ^d is from the center of all positive charges to the center of all negative charges in the donor. Similarly, the acceptor dipole is situated at the center of the acceptor and is along the unit vector ^ a and has dipole moment~ p A ¼ qe DA ^a, where DA ^a is a vector sum of fluctuating vectors oscillating at 3) In reality the nuclei do oscillate, but at frequencies that are irrelevant for FRET. j35 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.9 Ideal Dipole Approximation j 3 F€orster Theory Figure 3.5 A cartoon of a donor molecule, on the left, and an acceptor molecule, on the right. In reality, both donor and acceptor contain many charges, which vibrate and oscillate in several directions and at a range of frequencies on the order of 1 million GHz. When the size of the donor is much smaller than the donor– acceptor distance and the size of the acceptor is also much smaller than this distance, the ideal dipole approximation is valid. a range of frequencies. It points from the center of all positive charges to the center of all negative charges in the acceptor. We must realize that when a donor is excited by an electromagnetic wave at a certain frequency, the donor will start oscillating at a range of frequencies and not only at the frequency of the wave, because the oscillators that make up the donor are coupled. F€orster used the cgs unit system, which was the system of choice in the 1940s and 1950s. Today (2012) this system is hardly ever used, and in this presentation of F€ orster theory, the mks system is used, which is also known as the SI system. In the cgs system, the unit of charge is the statcoulomb (which is equal to the esu); but in the mks system, the unit of charge is the coulomb (C ¼ A s). One statcoulomb is equal to 3.3356 1010 C [33]. A consequence of this difference in units is that many equations in electromagnetism using the cgs system differ from the corresponding equations in the mks system [33]. However, the final equations derived by F€ orster do not depend on the system of units, but intermediate equations in the theory, for example, those for energy, power, and intensity, have different forms in the two systems. 3.10 Resonance as an All-or-Nothing Effect Resonance occurs when an oscillator capable of vibrating at a natural frequency interacts with an external system that forces this oscillator to vibrate at an external Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 36 frequency. The oscillator will pick up significant energy only if the external frequency and the natural frequency are equal or nearly equal. Resonance is everywhere: accomplished singers can break a wine glass by hitting the right note [34], a sturdy bridge can collapse when jolted at the right frequency [35], and most adults walk at a frequency of about 1 Hz because adult legs swing at a natural frequency of about 1 Hz. The resonance phenomenon that concerns us here occurs when acceptor charges oscillate in phase with donor charges at a distance. In his classical theory of energy transfer [32], F€ orster visualized a donor or an acceptor molecule as a group of coupled electrical oscillators. When the donor oscillators are in an external electromagnetic field caused by light that can excite the donor, the oscillators in the donor will oscillate and cause their own electromagnetic field. Acceptor oscillators often will not respond to the external electromagnetic field, but may be sensitive to the electric field from the donor oscillators. Let us first focus on one electrical oscillator inside an acceptor responding to an electric field generated by a donor at some distance away. We will call the direction in which the acceptor oscillator can swing the acceptor direction. The natural frequency of this oscillator is nACCEPTOR . The electric field is also along a certain direction, the donor field direction, which is not necessarily the same as the acceptor direction. However, this electric field must have a component along the acceptor direction, otherwise there will be no response. The amplitude of this component is E DF (which is equal to the amplitude of the donor field times the cosine of the angle between the two directions). The frequency of the donor field is nDONOR (which is also equal to the frequency of the donor oscillator). The electric field starts at time 0 and lasts for a certain amount of time t, which can vary. The energy W A that the acceptor dipole has at time t as a result of its interaction with the donor field depends on the frequency difference nACCEPTOR nDONOR . As a function of frequency, this energy has a strong maximum when this frequency difference is zero, but also has weak secondary maxima at other values. F€orster made the approximation to replace this intricate resonance behavior by a sharp rectangular peak. In other words, he assumed that either there is resonance or there is nothing: 8 > > 0; > > > < 2 2 qe E D WA ¼ t2 ; > 8me > > > > : 0; if if if nDONOR nACCEPTOR < 1 ; 2t 1 1 nDONOR nACCEPTOR ; 2t 2t 1 nDONOR nACCEPTOR > : 2t ð3:16Þ This approximation is schematically illustrated in Figure 3.6. From Equation 3.16 we see that the value of this peak increases drastically with time, but the width of the peak decreases with time. For example, doubling the time yields a fourfold increase in the value at the peak, but leads to a reduction by a factor of 2 in the width, as illustrated in Figure 3.7. j37 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.10 Resonance as an All-or-Nothing Effect j 3 F€orster Theory Figure 3.6 F€ orster replaced the resonance curve by a rectangular peak. Vertically the energy of the acceptor is plotted and horizontally the frequency. The center of the peak (for the curve as well as the rectangular peak) corresponds to the donor frequency being equal to the acceptor frequency. Figure 3.7 The height of the resonance peak is proportional to time-squared, but the width is proportional to the time. As a result, doubling the time yields a fourfold increase in the height of the peak, but leads to a reduction by a factor of 2 in the width. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 38 3.11 Details About the All-or-Nothing Approximation of Resonance The reader who has no problem accepting F€orster’s all-or-nothing approximation of resonance and is not interested in the reasons why this approximation is good should skip this section. It turns out that F€orster’s all-or-nothing approximation of resonance is excellent. It is not immediately obvious why. To prove that it is indeed very good, we must set up an equation of motion for the acceptor oscillator, solve it, calculate the energy of the acceptor, and compare this rigorous expression of the energy with the approximate equation displayed in Equation 3.16. Consider the forces acting on each acceptor oscillator. An electrical oscillator in the acceptor can oscillate along a certain line, the acceptor direction, which we will identify as the x-axis. The natural frequency of this oscillator is nACCEPTOR . This frequency is related to the spring constant k and the mass of the oscillator me by sffiffiffiffiffiffi 1 k nACCEPTOR ¼ : ð3:17Þ 2p me The x-component of the electric field generated by a donor oscillator has amplitude E D and frequency nDONOR and at time t is x-component of the electric field generated by donor ¼ E D cos ð2pnDONOR tÞ: ð3:18Þ From Equation 3.15, we know E D is equal to ED ¼ qe DD k : 4pe0 n2 r 3DA ð3:19Þ According to Newton’s second law, the net force on the oscillating charge, the sum of the elastic force kx and the electric force Q e E D cos ð2pnDONOR tÞ, equals the mass times the acceleration: me d2 x ¼ kx þ qe DD cos ð2pnDONOR tÞ; dt2 ð3:20Þ where x is the displacement and qe is the charge of the oscillator. Substituting k ¼ 4p2 me n2ACCEPTOR (from Equation 3.17) into (3.20), dividing by me, and utilizing the abbreviations u ¼ 2pnACCEPTOR and w ¼ 2pnDONOR transform (3.20) into d2 q DD x ¼ u2 x þ e cos ðwtÞ: dt2 me ð3:21Þ The solution of Equation (3.21) for the case where the initial displacement and the initial velocity are zero is x¼ qe E D ðcos ðwtÞ cos ðutÞÞ; me ðu2 w 2 Þ ð3:22Þ j39 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.11 Details About the All-or-Nothing Approximation of Resonance j 3 F€orster Theory so that the speed of the oscillator motion is dx uqe E D w sin ðutÞ sin ðwtÞ : ¼ dt me ðu2 w2 Þ u ð3:23Þ The energy of the acceptor, W A , the sum of the potential energy, ð1=2Þkx2 ¼ ð1=2Þme u2 x 2 , and the kinetic energy, ð1=2Þme ðdx=dtÞ2 , is " # 2 q2e E 2D 4u2 w 2 2 ð cos ð wt Þ cos ð ut Þ Þ : WA ¼ t þ sin ð wt Þ sin ð ut Þ 8me u ðw 2 u2 Þ2 t2 ð3:24Þ 2 For w ¼ u e with e p=t, the term in square brackets equals 1=ðetÞ and the term in curly brackets equals ðetÞ2 . Therefore, at very small differences between the donor and acceptor frequencies, this energy is equal to its maximum value ¼ W A;PEAK ¼ ½ðq2e E 2D Þ=8me t2 . Relevant values of t are on the order of the lifetime of the excited state, so that 2p=t is on the order of 1–10 GHz, whereas u and w correspond to frequencies in the UV or visible and are, therefore, on the order of a million gigahertz. As a result, W A has a series of minima that are essentially equal to zero for w ¼ u Np=t, where N is an even integer larger than 0. The case w ¼ u p=t corresponds to the border of the rectangular F€ orster peak. At that frequency, the actual W A value is about 0:4 W A;PEAK . With w ¼ u Np=t where N is an odd integer larger than 1, W A has secondary maxima equal to about 0:4 W A;PEAK =N 2 . The width of the all-or-nothing peak equals 1=t. The width of the actual resonance curve can be defined in terms of the area under the curve. The total area under the all-or-nothing peak and that under the actual resonance curve turn out to be essentially the same. Relevant intervals for the resonance curve are as follows: Width of frequency interval Frequency interval in terms of vACCEPTOR vDONOR Approximate area under the curve (% of total) 1/t 1/(2t) vACCEPTOR vDONOR 1/(2t) 77% 20/t 10/t vACCEPTOR vDONOR 10 99% Most of the energy is transferred near the end of the lifetime of the excited state of the donor, where t is on the order of nanoseconds, so that the width of the 99% interval is about 20 GHz. However, the relevant frequency values in the spectra are on the order of millions of gigahertz. This means that over 20 GHz, the spectra do not vary at all. Since the all-or-nothing peak and the actual resonance peak have the same area under the curve, the total amount of energy transferred in the all-ornothing approximation is equal to that transferred according to the actual resonance curve when t is on the order of a nanosecond. The all-or-nothing approximation can only fail if there is significant spectral variation over the 99% interval. Such variation is expected when t is very small, less than a femtosecond. However, early in the Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 40 process, the energy is extremely small because of the time-squared factor in the magnitude of the peak (energy after a femtosecond ¼ 1012 energy after a nanosecond). We conclude, therefore, that F€orster’s all-or-nothing approximation, replacing Equation 3.24 by Equation 3.16, is excellent. 3.12 Classical Theory Completed From Equation (3.16) we know that if one donor oscillator and one acceptor oscillator have the exact same frequency n ¼ nDONOR, the acceptor has, due to a resonance interaction with the donor, t s after the donor has started oscillating, an amount of energy W A equal to q2e E 2D t2 : 8me WA ¼ ð3:25Þ Expressing W A in terms of the donor–acceptor distance using Equation 3.19, we find q4e D2D k2 t2 WA ¼ 8me ð4pe0 Þ2 n4 r 6DA : ð3:26Þ D2D is proportional to W D , the energy of the donor oscillator, because 1 W D ¼ kD2D ¼ 2p2 me n2 D2D ; 2 ð3:27Þ meaning that D2D ¼ WD : 2p2 me n2 ð3:28Þ Substituting this into Equation 3.26 yields WA ¼ q4e k2 W D t2 16p2 me n2 ð4pe0 Þ2 n4 r 6DA : ð3:29Þ Now we must generalize this to the case where there is not just one frequency but distributions of frequencies for donor and acceptor. Such distributions can be described using oscillator strengths f eD ¼ f eD ðnÞ for the donor and f aA ¼ f aA ðnÞ for the acceptor. Specifically, f eD ¼ f eD ðnÞ ¼ probability to find a frequency between n and n þ dn: ð3:30Þ Remember that the width of the resonance peak, using F€ orster’s all-or-nothing approximation, is 1=t (see Equation 3.16). Therefore, relating Equation 3.30 to the corresponding acceptor frequency interval, we find f aA ðnÞð1=tÞ ¼ acceptor frequencies resonating with donor: ð3:31Þ j41 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.12 Classical Theory Completed j 3 F€orster Theory It follows, therefore, that we should multiply both sides of Equation 3.29 with f eD ðnÞf aA ðnÞð1=tÞdn and integrate over all frequencies: WA ¼ 1 ð q4e k2 W D t 16p2 me ð4pe0 Þ2 n4 r 6DA 0 f eD ðnÞf aA ðnÞdn : n2 ð3:32Þ Differentiating this with respect to time gives us F€ orster’s rate equation: dW A ¼ kT W D ; dt ð3:33Þ with kT ¼ 1 ð q4e k2 16p2 m2e ð4pe0 Þ2 n2 r 6DA 0 f eD ðnÞf aA ðnÞdn : n2 ð3:34Þ Substituting Equations 3.42 and 3.46 into Equation 3.34 yields kT ¼ 9ðln 10Þk2 WD 128p5 N 0 tD r 6DA 1 ð c4 f D ðnÞeA ðnÞdn : v4 ð3:35Þ 0 Using the definitions of the F€orster distance and the overlap integral, we find R60 ¼ 9ðln 10Þk2 WD J ; 128p5 n4 N 0 ð3:36Þ and this equation is identical to Equation 3.3, because N 0 ¼ N A . In (3.36) J stands for 1 ð J¼c 4 0 f D ðnÞeA ðnÞ dn; n4 which is Equation 3.8. And, we arrive at F€ orster’s equation: 6 1 R0 ; kT ¼ tD r 6DA ð3:37Þ ð3:38Þ which is Equation 3.2. Note that in this section the integrals extend from zero to infinite frequency as they are based on the theoretical model of coupled charged oscillators. However, integrals based on experimentally obtained spectra should only refer to the relevant part of the spectra as discussed in Section 3.5. 3.13 Oscillator Strength–Emission Spectrum Relation for the Donor Consider a donor molecule capable of emitting light. The electromagnetic energy this donor has is W D . If the quantum yield is WD , then the energy available for fluorescence is WD W D . This can be emitted over a range of frequencies. Ð1 The normalized fluorescence spectrum is f D ¼ f D ðnÞ, so that 0 f D ðnÞdn ¼ 1. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 42 The energy available for emission between n and n þ dn is equal to WD W D f D ðnÞdn. The average lifetime of the excited state is tD . Therefore, the rate of emission between n and n þ dn ¼ WD W D f D ðnÞdn : tD ð3:39Þ On the other hand, one electric oscillator having energy W D, mass me , charge qe , and frequency n, embedded in a medium with refractive index n, must radiate energy in accordance with Maxwell’s theory of electromagnetism, at a following rate: the rate of emission of one oscillator ¼ 8p2 nn2 q2e W D : 3ð4pe0 Þme c3 ð3:40Þ We have a distribution of oscillators with a fraction of f eD ðnÞdn between n and n þ dn. As a result, we have the rate of emission between v and v þ dv ¼ 8p2 nn2 q2e W D f eD ðnÞ : 3ð4pe0 Þme c3 ð3:41Þ Combining Equations 3.39 and 3.41 yields ½WD W D f D ðnÞ =tD ¼ ½8p2 nn2 q2e W D f eD ðnÞ = ½3ð4pe0 Þme c 3 Therefore, the relation between the donor oscillator strength and the donor emission spectrum is f eD ðnÞ ¼ 3ð4pe0 Þme c3 WD f D ðnÞ : 8p2 nn2 q2e tD ð3:42Þ 3.14 Oscillator Strength–Absorption Spectrum Relation for the Acceptor Imagine electromagnetic radiation falling upon 1 cm2 of a layer containing acceptor molecules at a concentration c A moles=l. This layer has a very small thickness of ‘ cm. Consider the spectral energy density s ðnÞ, defined such that s ðnÞdn represents the electromagnetic energy per unit of volume in the frequency range between n and n þ dn. From the Lambert–Beer law, we find that the spectral energy density absorbed in this layer is equal to the transmitted minus the incident spectral density, that is, absorbed spectral energy density ¼ s ðnÞ 1 eeA cA ‘ , where eA ¼ eA ðnÞ is the molar extinction coefficient of the acceptor in units 1/(cm M). Because ‘ is very small, and therefore eA c A ‘ is very small, the following simplification is valid: h i s ðnÞ 1 10eA cA ‘ ¼ s ðnÞ 1 eðln 10ÞeA cA ‘ ¼ s ðnÞ½1 f1 ðln 10ÞeA cA ‘g ¼ s ðnÞðln 10ÞeA cA ‘: Here, ln 10 is the natural logarithm of 10 (see Section 3.3). Therefore, ðln 10Þs ðnÞeA c A ‘ is the spectral energy density per unit of volume absorbed. This energy is absorbed in an extremely short time. The speed at which this radiation propagates in the medium is c=n (c is the speed of light in vacuo and n j43 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.14 Oscillator Strength–Absorption Spectrum Relation for the Acceptor j 3 F€orster Theory is the refractive index of the medium). Therefore, if we are interested in the energy absorbed in 1 s, we must visualize a cylinder with a length of c=n and a crosssectional area of 1 cm2 just in front of the layer. The energy distributed over this cylinder is the energy per second entering the layer. So, the energy per second entering the layer is sðnÞ c=n, and the energy absorbed in it is ðc=nÞðln 10Þs ðnÞeA cA ‘. This layer has a volume of ‘ 1 cm3 ¼ ‘ 1 ml. Since the concentration of acceptor is c A moles=l, this volume contains ‘ cA millimoles, which is N 0 ‘cA acceptor molecules (N 0 is the number of molecules per millimole) (see Section 3.3). Thus, energy absorbed per second and per molecule ¼ ðln 10ÞcsðnÞeA : nN 0 ð3:43Þ On the other hand, from classical electromagnetic theory, we know that one electric oscillator with mass me , charge qe , and frequency n, embedded in a medium with refractive index n, where the spectral energy density is sðnÞ will absorb energy at a predictable rate, the energy absorbed per second by one oscillator ¼ pq2e sðnÞ : 3n2 ð4pe0 Þme ð3:44Þ We have a distribution of oscillators with a fraction of f aA ðnÞdn between n and n þ dn. As a result, we have the energy absorbed per second per molecule ¼ pq2e sðnÞf aA ðnÞ : 3n2 ð4pe0 Þme ð3:45Þ Combining Equations 3.43 and 3.45 yields ðln 10ÞceA ðnÞs ðnÞ pq2e s ðnÞf aA ðnÞ ¼ : nN 0 3n2 ð4pe0 Þme Therefore, the relation between acceptor oscillator strength and acceptor extinction spectrum is f aA ðnÞ ¼ 3ðln 10Þð4pe0 Þnme ceA ðnÞ : pN 0 q2e ð3:46Þ 3.15 Quantum Mechanical Theory When charges are bound to each other inside a molecule, the energies available to them do not form a continuous spectrum, but the energy values are quantized. Resonance energy transfer can, therefore, be understood as coupled transitions, as shown in Figure 3.8. This diagram is essentially the same as the energy level diagram introduced by F€orster in his 1959 paper [3]. According to quantum mechanics, a system can adopt a number of different states. Considering a donor and an acceptor, it is clear that transfer may take place Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 44 Figure 3.8 Simplified energy-level diagram of resonance energy transfer. D refers to the donor and A to the acceptor; asterisks denote excited states. (Adapted from Ref. [2].) from the state in which the donor is excited and the acceptor is not, jD Ai, to a state in which the acceptor is excited and the donor is not, jDA i. The interaction responsible for transfer at donor–acceptor distances larger than the sizes of the charge distributions is the dipole–dipole interaction U, which is given by U¼ 1 ½~ p A 3ð~ p D ^r Þð^r ~ p AÞ ; p D ~ 4pe0 n2 r 3DA ð3:47Þ where n is the refractive index of the medium, r DA denotes the distance between the donor and the acceptor, ^r is a unit vector pointing from the donor to the acceptor,~ pD and~ p A are the dipole moment vectors of the donor and acceptor charge distribution, respectively, ~ p D ~ p A is the dot product of these two vectors, that is, the projection of one on the other. According to the time-resolved perturbation theory, the rate of transfer in the “very weak coupling” [5] limit is ð 1 kT ðW D ; W A Þ ¼ ð3:48Þ hD AjU jDA i2 dW; h where kT ðW D ; W A Þ is the rate of transfer from an excited donor molecule with initial W D to an acceptor with initial energy W A, h is Planck’s constant, and the integral is over all possible values of the transferred energy W. (The meaning of “very weak”, “weak”, and “strong coupling” in this context is discussed by F€ orster [5], Kasha [36], and Knox [20]. It is safe to assume that the Born–Oppenheimer approximation is valid. This approximation states that the electronic motion and j45 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.15 Quantum Mechanical Theory j 3 F€orster Theory the nuclear motion in molecules can be separated, so that the wave function for the molecule can be considered to be a product of an electronic wave function times a nuclear wave function. As a result, the expectation value for the energy in Equa~ D and tion 3.48 can be expressed in terms of electronic transition dipole moments m ~ A , of the donor and acceptor, respectively. These vectors can be written as m ~ D ¼ jmD j^d and m ~ A ¼ jmA j^a; m ð3:49Þ where ^d and ^a are unit vectors in the direction of the donor and acceptor transition moments, respectively, and, jmD j and jmA j represent the magnitudes of these moments. In this approximation, the integrand in Equation 3.48 can be expressed in terms of the electronic transition moments as follows: hD AjU jDA i2 ¼ k2 m2D m2A ð4pe0 Þ2 n4 r 6DA S2D S2A ; ð3:50Þ where k2 is the orientation factor defined in Section 3.7 and m2D m2A is the square of jmD jðjmA jÞ. The factors SD and SA represent vibrational overlap integrals: SD ¼ SD W D ; W D W is the overlap integral between the initial vibrational donor state with energy W D and the final state with energy W D W, and SA ¼ SA W A ; W A þ W is the overlap integral between the initial vibrational acceptor state with energy W A and the final state with energy W A þ W. From the transfer rate kT ðW D ; W A Þ of Equation 14.2, we can obtain the total transfer rate of thermal equilibrium by introducing suitable Boltzmann factors and integrating over all energies W D and W A . These Boltzmann factors g W D for the excited donor and g ðW A Þ for the acceptor in the ground state are continuous functions and are normalized on an energy scale. Therefore, by multiplying both sides of Equation 3.48 by g W D dW D and g ðW A ÞdW A , integrating over all W D and W A , and changing integration variable from energy to frequency: W to n, we obtain the following expression for the total transfer rate kT : kT ¼ with and k2 ð4pe0 Þ2 n4 h2 r 6DA 1 ð M D ðnÞLA ðnÞdn; ð3:51Þ 0 ð MD ðnÞ ¼ m2D g W D S2D W D ; W D hn dW D ð3:52aÞ ð LA ðnÞ ¼ m2A g ðW A ÞS2A ðW A ; W A þ hnÞdW A : ð3:52bÞ Analyses similar to that done in Sections 3.12 and 3.13 [32] show that M D ðnÞ is related to the normalized fluorescence spectrum of the donor and LA ðnÞ is proportional to the extinction spectrum of the acceptor: MD ðnÞ ¼ ð4pe0 Þ3hWD c3 f D ðnÞ : 32p3 ntD n3 ð3:53aÞ Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 46 LA ðnÞ ¼ ð4pe0 Þ3ðln 10ÞnhceA ðnÞ : 4p2 N 0 n ð3:53bÞ Substituting Equations 3.53a and 3.53b into Equation 3.51 yields F€ orster’s equation for the rate of transfer: kT ¼ 9ðln 10Þk2 Q D 128p5 N 0 tD r 6DA 1 ð c4 f D ðnÞeA ðnÞdn : v4 ð3:54Þ 0 This equation is the same as Equation 3.35. As noted in Section 3.12, the theoretical integrals extend from zero to infinite frequency, but integrals based on experimentally obtained spectra should only refer to the relevant part of the spectra as discussed in Section 3.5. 3.16 Transfer in a Random System Consider an ensemble of donor and acceptor molecules, belonging to different species, for which the following assumptions hold: 1) The molecules are randomly distributed through three-dimensional space. 2) Resonance energy transfer is possible at an appreciable rate from donor to acceptor, but transfer in the opposite direction is negligible. 3) Translational diffusion is slow compared to the rate of transfer, so that the distances between donor–acceptor pairs do not change significantly during the time transfer takes place. F€orster has pointed out [3] that these conditions are approximately met in solutions of moderated viscosity, in which case the Brownian rotational motion for both donor and acceptor molecules is also much faster than the transfer and is unrestricted, so that the orientation factor can be set equal to 2/3. However, in many biological systems, these assumptions may not be correct (see Chapter 4). In a system in which these assumptions apply, consider a donor molecule that is already excited at time t ¼ 0. If no acceptor molecules had been present, it would lose its excitation energy after an average lifetime tD through radiation or nonradiative deactivation processes. Its natural rate of deactivation is, therefore, 1=tD . The presence of an acceptor molecule at a distance r k provides another deactivation pathway for the excited donor molecule. The rate of transfer from the donor to the acceptor is, according to F€orster theory, ð1=tD ÞðR0 =r k Þ6 . Because of these two competing processes, the probability r ¼ rðtÞ that the donor molecule is still excited at time t is given by the following equation: " # N d 1 X R0 6 r; ð3:55Þ þ r¼ dt tD k¼1 r k where the summation is over all N acceptor molecules in a spherical volume around the excited donor molecule with a radius much larger than the F€ orster distance R0 . j47 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.16 Transfer in a Random System j 3 F€orster Theory The solution of this differential equation with the condition that rð0Þ ¼ 1 is ( " # ) N 1 X R0 6 t : ð3:56Þ rðtÞ ¼ exp þ tD k¼1 r k tD The donor fluorescence is proportional to the average of this quantity, hrðtÞi. F€orster [4] showed that this average can be written as in Equation 3.57. The derivation of Equation 3.57 is explained in Section 3.17, where other details about transfer in a random system are also given: pffiffi rðtÞ ¼ exp s 2ðc=c0U Þ s ; ð3:57Þ where s ¼ t=tD and c 0U is the “critical concentration for heterotransfer,” which is given by 3 3 c0U ¼ pffiffiffiffiffi 0 3 ¼ pffiffiffiffiffi ; 2 p3 N R0 2 p3 N A R30 ð3:58Þ Note that there are two critical concentrations: c 0U and c 0L . c 0U is the critical concentration for heterotransfer, that is, transfer between unlike molecules (U stands for unlike), and c 0L is the critical concentration for homotransfer, that is, transfer between like molecules (L stands for like), discussed in Section 3.17. The efficiency E can also be calculated (see Section 3.17) and is given by pffiffiffi E ¼ px ðex Þ2 f1 erf ðx Þg; with x ¼ c=c0U ; ð3:59Þ where erf is the error function, which is defined below, in Equation 3.76. The efficiency is plotted in Figure 3.9 versus x ¼ c=c 0U . It turns out that when c ¼ c 0U , the efficiency is equal to 76%. 3.17 Details for Transfer in a Random System The average of rðtÞ, defined in Equation 3.56, plays a key role in F€ orster’s theory of heterotransfer in a random system of donors and acceptors [4]. This average can be written as hrðtÞi ¼ et=tD ½HðtÞ N ; ð3:60Þ with RðV H ð tÞ ¼ 6 eðR=R0 Þ t=tD w ðRÞdR; ð3:61Þ 0 where RV is the radius of the sphere around the excited donor molecule that contains the N acceptor molecules to which transfer can occur, w ðRÞdR represents the probability for finding an acceptor molecule at a distance between R and R þ dR from the excited donor molecule. The assumption of randomness (the first Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 48 Figure 3.9 The efficiency versus concentration for a system of random donors and acceptors where the rotational diffusion is fast, but the translational diffusion is slow compared to the rate of transfer. Here, C is the acceptor concentration and C0 is the critical concentration for heterotransfer, C0 ¼ C0U, defined in Equation 3.58. The graph follows from Equation 3.59. assumption in Section 3.15) implies that w ðRÞ is such that every point in the sphere has equal probability for being occupied. Therefore, w ðRÞ must be equal to w ðRÞ ¼ 3R2 =R3V : ð3:62Þ Substituting this into Equation 3.61 allows us to transform HðtÞ into 1 pffiffiffiffiffi HðtÞ ¼ zV 2 1 ð zV ez pffiffiffiffiffi dz; z3 ð3:63Þ where z and zV are defined as 6 R t z¼ ; R 0 tD 6 R0 t : zV ¼ RV tD ð3:64Þ ð3:65Þ Through integration by parts, we obtain 1 ð zV ez 2ezV pffiffiffiffiffi dz ¼ pffiffiffiffiffi 2 zV z3 1 ð 0 ez pffiffiffiffiffi dz þ 2 z3 zðV 0 ez pffiffiffiffiffi dz: z3 ð3:66Þ Note that for all relevant values of the time t, zV is much smaller than 1, because RV is assumed to be much larger than R0 . Therefore, it is a very good approximation pffiffiffiffiffi to expand the right-hand side of Equation 3.66 in powers of zV and to keep only the j49 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.17 Details for Transfer in a Random System j 3 F€orster Theory first two terms. Applying this approximation and substituting Equation 3.66 into Equation 3.63 yields pffiffiffiffiffiffiffiffi pzV : H ð tÞ ¼ 1 ð3:67Þ As a result, the average of rðtÞ in Equation 3.60 can be rewritten as hrðtÞi ¼ et=tD ð1 pffiffiffiffiffiffiffiffi N pzV Þ : ð3:68Þ Because the number N can be assumed to be extremely large, Equation 3.68 can be simplified to N pffiffiffiffiffiffi 1 pzV ¼ et=tD N pzV : ð3:69Þ hrðtÞi ¼ et=tD lim 1 N pffiffiffiffiffiffi N!1 N Employing the definition of zV and introducing x defined by x¼ pffiffiffi 3 pNR0 ; 2R3V ð3:70Þ the average of rðtÞ becomes pffi hrðtÞi ¼ es2x s ; ð3:71Þ where s ¼ t=tD . The quantum yield in the presence of acceptor, WDA , is 1 ð ~ WDA ¼ C hrðtÞidt; ð3:72Þ 0 ~ is a constant. The quantum yield in the absence of acceptor, WD , can be where C calculated from Equations 3.72 and 3.69 for N ¼ 0: ~ WD ¼ C 1 ð ~ D: et=tD dt ¼ Ct ð3:73Þ 0 Combining Equations 3.72 and 3.73 yields WDA ¼ WD where y ¼ 1 ð pffi s2x s e 0 1 ð ds ¼ ex 2 y2 ð3:74Þ ds; 0 pffiffi s þ x, so that s ¼ ðy x Þ2 and ds ¼ 2ðy xÞdðy x Þ. Therefore, WDA ¼2 WD 1 ð x 2 y2 e 0 x2 1 ð ðy xÞdðy x Þ ¼ 2e y2 e 0 pffiffiffi 2 ydy xe p pffiffiffi p x2 1 ð ey dy: 2 x ð3:75Þ The first integral on the right-hand side equals 1 and the second can be expressed in terms of the error function, which is defined as Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 50 1 ð 0 x 2 2 2 erf ðxÞ ¼ pffiffiffi ey dy ¼ 1 pffiffiffi p p ey dy: 2 ð3:76Þ Combining Equations 3.75 and 3.76 allows us to express the efficiency in terms of x: E ¼1 WDA pffiffiffi x2 ¼ pxe f1 erf ðxÞg: WD ð3:77Þ Note that if we choose N ¼ N A in Equations 3.60 and 3.70, the concentration becomes c ¼ 3= 4pR3V . As a result, x in Equation 3.70 is equal to pffiffiffiffiffi x ¼ 2=3 p3 N A R30 c; ð3:78Þ and because of the definition of c 0 ¼ c 0U in Equation 3.58, we see that x ¼ c=c 0 , confirming Equation 3.59. 3.18 Concentration Depolarization Concentration depolarization is a homotransfer phenomenon. In a system where the fluorophores belong to a single species, FRET results in a strong depolarization of the fluorescence. The excitation energy of a molecule that absorbs a photon at a certain moment may jump from molecule to molecule until emission occurs at a later time. Thus, a fluorescence photon, which in dilute solution is emitted by the absorbing molecule, may in concentrated solutions be emitted by another molecule. This process does not affect the time dependence of the fluorescence intensity, but it broadens the angular distribution of the emission transition moments and consequently gives rise to depolarization of the emission. A graph of the fluorescence polarization versus the logarithm of the concentration shows a constant level of depolarization at low concentrations and a sharp drop at higher concentrations [2]. In his theory of concentration depolarization, F€orster assumed that only the photons emitted by the primary molecule are maximally polarized and that the other photons are completely unpolarized [2]. He derives the polarization as a function of the concentration c for c c0 and for c c 0, where c 0 ¼ c0L , that is, the critical concentration for homotransfer: c0L ¼ 3 3 ¼ : 4pN 0 R30 4pN A R30 ð3:79Þ The depolarization depends on p1 , the probability that the fluorescence is emitted by the initially excited molecule. If depolarization is due to concentration quenching only and rotational motion can be ignored, p1 is equal to r=r 0 , the ratio of the anisotropy and the fundamental anisotropy (¼ anisotropy in the absence of motion or transfer). In the low concentration limit, only the interaction between the primary molecule and one other is considered, and p1 is given by 1 ð p1 ¼ 0 1 þ t kT j e dj; 1 þ 2t kT c c0 ; ð3:80Þ j51 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.18 Concentration Depolarization ðx j 3 F€orster Theory where t is the fluorescence lifetime in the absence of transfer, kT is the rate of transfer, and j ¼ 0:001 N A cV, where N A ¼ Avogadro’s number, c ¼ concentration in moles/l, and V ¼ 4pR3 =3– here R denotes the distance between molecules. In the high concentration limit, the excitation energy is thought to “diffuse” away from the primary molecule and p1 is given by p1 ¼ 1 ð1 þ 1:55=cÞe1:55=c ; c c0 : ð3:81Þ pffiffiffiffiffiffiffiffiffiffiffiffiffi Here c ¼ ðc=c 0 Þ W=W0 , where W is the quantum yield and W0 is the quantum yield in the absence of transfer. This theory has been further developed by Knox and others (see Section 3.19). 3.19 FRET Theory 1965–2012 F€orster’s work inspired an enormous volume of both experimental and theoretical work, not to mention applications and patents. The concluding section of this chapter is an overview of the theoretical work inspired by his results. However, this overview does not include work on kappa-squared. This has been dealt with in Chapter 4. The theory of concentration depolarization and quenching has been further developed by Knox, Craver, and others [36–44] (see Refs [45,46] for reviews). Craver and Knox compared different theories for concentration quenching in three dimensions and showed that experimental data were in good agreement with their extension of F€orster’s theory [39]. Craver [40] has proposed a “universal” curve for concentration depolarization in three dimensions. This curve, which is shown in Figure 3.10, fits experimental data quite well [40,47]. The GAF theory deals with the time dependence of transport of electronic excitation between like molecules that are randomly distributed [44]. The theory predicts the time dependence of GS ðtÞ, the probability of finding the excitation on the initial site as a function of the time t after excitation. This function can be observed in a picosecond transient (holographic) grating experiment. In this experiment, a delayed picosecond probe pulse is Bragg diffracted by a grating that is optically produced in the sample by the interference of two coherent picosecond excitation pulses. Absorption by the sample in the overlap region of the two excitation pulses results in a spatially varying sinusoidal distribution of excited states resulting in Bragg diffraction of the probe pulse. The intensity of the diffracted probe pulse is proportional to the square of the difference in the absorption between the grating peaks and nulls. Time-dependent processes that reduce this peak–null difference result in the decay of the diffracted signal [44,48]. In the GAF theory, the Laplace–Fourier transform of GS ðtÞ is expanded as a diagrammatic series. Topological reduction of the series establishes an analogy of diffusion. This diagrammatic technique also suggests an interesting class of self-consistent approximations. One of these self-consistent approximations is applied to the specific case of the F€ orster Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 52 Figure 3.10 The “universal curve” proposed by Craver for concentration depolarization. The ratio r=r 0 , the steady-state anisotropy over that in the p absence ffiffiffiffiffiffiffiffiffiffiffiffi of transfer, is plotted versus x 0 ¼ 2 FD =pðc=c0 Þ, where FD is the quantum yield in the absence of transfer, c is the concentration, and c0 ¼ c0L is the critical concentration for homotransfer, defined in Equation 3.79. This curve is in good agreement with experimental data [40,47]. transfer rate. The solutions obtained are well behaved for all times and all site densities and indicate that transport is nondiffusive at short times, but diffusive at long times. The mean-squared displacement of the excitation and its time derivative are calculated. These calculations illustrate that the time regime in which diffusive transport occurs is dependent on density. In systems with low density, transport of electronic excitation becomes diffusive only at times longer than a few minutes; whereas for high densities, transport becomes diffusive within one lifetime of the excited state [44]. Baumann and Fayer have discussed excitation transfer in the disordered two-dimensional and anisotropic three-dimensional systems for the cases of heterotransfer (direct trapping) in two-component systems and homotransfer (donor–donor transfer) in one-component systems [43]. Using the twoparticle model proposed by Huber et al. [49], Baumann and Fayer calculate the configurational average of GS ðtÞ. For the isotropic three-dimensional case treated by Huber et al., excellent correspondence is found with the GAF theory. The anisotropy of the dipole–dipole interaction is included in the averaging procedure. Two regimes of orientational mobility are considered: the dynamic and static limit, rotations being much faster or slower, respectively, than the energy transfer. Several geometrical distributions are investigated. The fluorescence anisotropy decay, which can be studied in a transient grating experiment or in a florescence depolarization experiment, is a useful observable for GS ðtÞ in homotransfer [43]. Baumann and Fayer focus on nonradiative transport [43]. A unified treatment of radiative and nonradiative transport was introduced by Berberan-Santos et al. [50] and the role of radiative transport has been reviewed by the same authors [51]. Huber et al. [49] report on the time dependence of fluorescence line narrowing. In the system j53 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.19 FRET Theory 1965–2012 j 3 F€orster Theory studied, a background is observed around a narrow band of frequencies. The appearance of this background in the fluorescence spectra is indicative of the transfer of excitation from fluorophores inside the band to molecules whose optical frequencies lie outside the band. The authors treat back-transfer effects in a variety of approximations and compare their theory with experimental data [49]. Hart et al. [52] used time-correlated single-photon counting to measure GS ðtÞ by monitoring the fluorescence concentration depolarization for a dye in glycerol. Hart et al. found that the three-body GAF theory accurately describes the fluorescence depolarization at the lower dye concentrations. At higher concentrations, the measured GS ðtÞ was found to be perceptibly slower than that predicted by GAF theory. The authors suggest that this deviation may arise from nonrandom dye distributions in solution, rather than from errors in the three-body GAF theory. They note that the experimental decay can also be described at all concentrations by the Huber–Hamilton– Barnett model [49]. The review by Knox [37] is primarily concerned with the theory of excitation migration. However, he also discusses experimental data on the rate of pairwise excitation transfer between like molecules, with particular attention to chlorophyll a. In a more recent review [20], Knox notes that the CDA (see Equation 3.2b) for chlorophyll a is about 68 4 nm6/ps. Kawski’s review [45] on “excitation energy transfer and its manifestation in isotropic media” is thorough and lists a large number of relevant references. It discusses essentially all aspects of energy transfer. However, the paper devotes particular attention to the effect of excitation energy migration on fluorescence anisotropy [45]. Fluorescence depolarization due to homo- and hetero-FRET was analyzed by Berberan-Santos and Valeur [53] and reviewed in Ref. [54], which elegantly describes many other energy transfer phenomena [54]. The theory of FRET on surfaces and membranes has been an active field [55–64]. Of these references, the work by Wolber and Hudson [56] probably had the most impact. These authors have found an analytical solution of the FRET problem in two dimensions for the case where the orientation factor is independent of the donor– acceptor distance and both donors and acceptors are randomly distributed in a plane [56]. In Refs [55–64], the emphasis is on heterotransfer. Homotransfer allows studying the accumulation of proteins in membranes. The theoretical framework that relates fluorescence anisotropy to cluster size has been provided by Runnels and Scarlata [65], who employ a theoretical analysis of homotransfer in clusters of like molecules all containing the same fluorophore. In its simplest form, the Runnels– Scarlata theory predicts that the anisotropy of a cluster of N molecules equals the anisotropy of the monomer divided by N [65]. Towles et al. have applied Monte Carlo simulations to study microheterogeneity and domain size in membranes. They conclude that Monte Carlo calculations clearly indicate that FRET is indeed sensitive to domain sizes in the range of 5–50 nm, but that a specific model is required to obtain a value for the domain size [66]. The idea of tryptophan imaging of membrane proteins has been proposed and analyzed by Kleinfeld [67]: tryptophans in membrane proteins serve as donors and anthroyloxy fluorophores serve as acceptors with the anthroyloxy group attached to lipids at various distances from the midplane of Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 54 the membrane. Kleinfeld and Lukacovic successfully applied this idea to locate the tryptophan-109 in cytochrome b5 [68] confirming an earlier conclusion by Fleming et al. [69]. Other interesting membrane FRET studies are mentioned in Refs [70–75]. Fluorescence studies of membrane heterogeneity has been reviewed by Davenport [76]. A theory of FRET in micelles has been developed [22]. Photosynthesis motivated substantial theoretical work [77–79,104,105]. Yang et al. [80] point out that in the theory of energy transfer in photosynthesis, F€ orster’s ideas can be successfully applied, but that Redfield theory [81] is more appropriate when the Coulombic coupling is greater than the electron–phonon coupling strength. Actually, there are three levels of discourse to consider when reviewing the theory: (i) F€orster process, which is the transfer or delocalization of an initially localized excited state. (ii) F€orster theory, which is his selection of a definition of rate of transfer and a method to calculate it. (iii) F€orster’s equation itself, which is a result of his applying his theory to the dipole–dipole case [20]. Table 3.3, adapted from Scholes [28] with minor changes, presents a history of coupling models in FRET. Hauser et al. generalized F€orster’s equation for energy transfer in three dimensions to the case of one, two, or three dimensions [106]. Often it is necessary to consider excluded-volume effects due to the geometry of the system, which prevents the acceptor from penetrating a certain volume surrounding the donor. Such excluded-volume effects have been discussed by Blumen et al. [107], Wolber and Hudson [56], Duportail et al. [108], and Tcherkasskaya et al. [109]. The authors of Refs [108,109] made use of the stretched exponential model introduced by Drake et al. [110]. Dewey [111] has reviewed the relations between FRET and fractals. If the donor–acceptor distance can change because of the lateral diffusion during the excited-state lifetime of the donor, FRET can be enhanced [112–116]. The parameter determining the degree of this enhancement is Z ¼ DtD =s2 ; ð3:82Þ where D is the sum of the lateral diffusion coefficients of donor and acceptor, tD is the donor lifetime in the absence of transfer, and s is the mean donor–acceptor distance. Three regimes can be distinguished [113]: 1) Z 1, the static limit, where the transfer is low and constant, that is, there is essentially no variation with diffusion. 2) Z 1, the intermediate regime, where the efficiency is sensitive to diffusion. 3) Z 1, the rapid diffusion limit, where the efficiency approaches a maximum value and again becomes independent of diffusion. Since distances of interest are in the 1–10 nm range and the diffusion coefficient of a typical fluorophore in aqueous solvents is on the order of 106 cm2/s, the donor lifetime in the case of rapid diffusion should be several orders of magnitude above the conventional nanosecond range. This technique of rapid diffusion FRET is reviewed by Stryer et al. [117]. Kouyama et al. [118] and Mersol et al. [119] have discussed the effects of restricted rotation in diffusion-enhanced FRET. j55 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.19 FRET Theory 1965–2012 j 3 F€orster Theory Table 3.3 A history of coupling models in resonance energy transfer. Authors Application Comments Craig and Walmsley [82], Davydov [83], Kasha et al. [84], McClure [85] Exciton states F€ orster [2,5] Electronic energy transfer Buckingham and Dalgarno [86] The interaction of ground-and excitedstate helium atoms Dexter [25] and Merrifield [87] Triplet–triplet energy transfer Koutecky and Paldus [88,89] Transannular interactions Andrews [90], Avery [91], Craig and Thirunamachandran [92], McLone and Power [93], Scholes and Andrews [94] Azumi and McGlynn [95], Murrell and Tanaka [96] Very long-range coupling Dipole–dipole coupling is used to define the electronic states of molecular aggregates. This dipole–dipole interaction arises from Coulomb forces. Higher multipoles are ignored Development of a theory for the rate of energy transfer, through the dipole–dipole coupling Interaction is between a ground-state helium atom and a helium atom in the first triplet or singlet metastable state. Heitler– London method is used Orbital overlap effects considered to arise via an exchange integral obtained from the Coulombic integral by permutation of two orbitals Calculation of the interactions between close molecules and perturbations of their absorption spectra Quantum electrodynamical theories for the form of dipole–dipole coupling over very large distances, including the near field, the intermediate zone, and the far field Naqvi [97], Naqvi and Steel [98] Exchange-induced resonance energy transfer LMO coupling model Harcourt et al. [99], Scholes and Ghiggino [100], Scholes and Harcourt [101], Scholes et al. [102] Scholes et al. [103] Excimers Special cases in photosynthesis Calculation of spectra based on the LMO (localized molecular orbital) prescription involving locally excited and charge transfer configurations Theory based on the exchange interactions in singlet/triplet–singlet/doublet energy transfer Orbital overlap-dependent coupling (LMO model), revealing that the significant overlap-dependent coupling is mediated via charge transfer configurations A study of couplings involving the carotenoid S1 state If the parameter Z in Equation 3.82 is on the order of 1, the rate of translational motion is of the same order as the rate of transfer, and FRET can be employed to measure the lateral diffusion of the donor and/or acceptor [112,115,120–123] or from fluctuations in the FRET efficiency [124]. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 56 Tanaka and Mataga [125] studied theoretically the effects of internal rotation on the decay and the anisotropy of a donor and an acceptor bound to a spherical macromolecule. The model system they considered is one in which the donor is internally rotating around an axis fixed at the macromolecule and the acceptor has a fixed position and orientation [126]. The relevant parameter is kT =DR , the average rate of transfer over the rotational diffusion constant. If this parameter is small, both the donor fluorescence and the anisotropy are single exponentials. However, if this parameter increases, the deviation from single-exponential behavior becomes more and more pronounced [125]. In general, rotational motion when present in combination with FRET will affect the time dependence of both the fluorescence and the anisotropy. This does not mean, however, that time dependence in the fluorescence anisotropy must result from rotational motion. Energy transfer or other excited-state reactions can give rise to a strong time dependence in the anisotropy in the absence of rotational motion [126–128]. This relation with time is due to coupling between two states with different transition moments [126–128]. Van der Meer et al. [129] proposed a general method to take into account the effects of motion on FRET. This method is applicable to both rotational and translational motion and is based on the idea that a system exhibiting both motion and FRET can be modeled by specifying a number of “states” and the rates of transitions between them. A state in this context is a set of conditions and specific coordinates that describe the system at a certain moment in time. There are excited-donor states (in which the donor is excited, but not the acceptor), excited-acceptor states (in which the acceptor is excited, but not the donor), and the states without excitation (neither donor nor acceptor is excited). A transition from an excited-donor state to an excited-acceptor state represents energy transfer, whereas a transition from an excited-donor state to another excited-donor state with a different position or orientation portrays translational or rotational motion. Fluorescence corresponds to a transition from an excited state to a state without excitation. Photoselection determines the initial occupation values of the states. This method results in matrix equations that are linear differential equations in time. The time dependence of the intensities and anisotropies of donor and acceptor can be expressed in terms of eigenvectors and eigenvalues of matrices [129]. The simplest example of this approach is to have a donor–acceptor distance that can only be equal to R1 (short) or R2 (long) with the distance changing between these two values at a rate J, and both donor and acceptor having isotropically degenerate transition dipoles (or undergoing very fast rotations). For this model, the donor fluorescence at time t after excitation with a short pulse is I ¼ I 0 exp ððJtD þ 1Þt=tD Þ a exp ðt=tD ÞðR0 =R2 Þ6 Jt=p þ ð1 aÞ exp ðt=tD ÞðR0 =R1 Þ6 þ Jt=p ; h i 2JtD = ðR0 =R1 Þ6 ðR0 =R1 Þ6 ðp 1Þ2 rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi with p ¼ h i 2ffi and a ¼ 2ðp2 þ 1Þ ; 1 þ 1 þ 2JtD = ðR0 =R1 Þ6 ðR0 =R1 Þ6 ð3:83Þ j57 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 3.19 FRET Theory 1965–2012 j 3 F€orster Theory where I 0 is the initial intensity, tD is the donor lifetime in the absence of acceptor, and R0 is the F€orster distance. If 2JtD ðR0 =R1 Þ6 ðR0 =R2 Þ6 , the fluorescence intensity becomes h i 1 I ¼ I 0 exp ððJtD þ 1Þt=tD Þ exp ðt=tD ÞðR0 =R2 Þ6 þ exp ðt=tD ÞðR0 =R1 Þ6 ; 2 ð3:84Þ but for 2JtD ðR0 =R1 Þ6 ðR0 =R2 Þ6, one finds I ¼ I 0 exp JtD þ 1 þ ðR0 =R1 Þ6 t=tD ; ð3:85Þ which depends only on the distance of closest approach, as expected, and contains the FRET enhancement factor JtD. Canley et al. discussed FRET efficiency at high excitation intensity. They derived the following equation for the FRET efficiency: E¼ 1 L þ ðR=R0 Þ6 ; ð3:86Þ where L ¼ 1 represents the well-known weak field case. They showed that L can be significantly larger than 1. The case L > 1 reflects the inability of doubly excited dye pairs to undergo energy transfer [130]. Raicu has developed a theoretical model for FRET from a single donor to multiple acceptors and from multiple donors to a single acceptor [131]. Bojarski et al. studied the possibility of FRET from a single donor to multiple acceptors using Monte Carlo techniques [132]. Rolinski and Birch have introduced new ideas about donor–acceptor distributions [133] and lifetime distributions [134]. Swathi and Sebastian pointed out that for energy transfer from a dye to a nanotube, one can use the dipole approximation for the dye, but not for the nanotube [135]. Consistent with this finding is the conclusion by Wong et al. that the point dipole approximation is inappropriate for use with elongated systems such as carbon nanotubes and that methods that can account for the shape of the particle are more suitable [136]. In the metal-enhanced fluorescence, a new field with a vast potential for applications [137], resonance energy transfer plays a significant role [138]. Lakowicz introduced the radiative plasmon model and showed that this model is consistent with a wide range of experimental results, including FRET from fluorophores to nearby metal surfaces [138]. Acknowledgments I wish to thank Dr. Bob Knox for stimulating discussions, useful advice, and making me aware of papers I had overlooked. I am grateful to Dr. Joggi Wirz, who discovered that the 9000-form of the F€orster equation is incorrect while working on his book [139]. We had highly interesting correspondence about this topic. Thanks to Dr. David Andrews who shared with me his insights into QED&FRET and the Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 58 relevance of the spectral overlap. I am indebted to Dr. Manuel Prieto for sending me relevant papers and giving me useful suggestions. I am also grateful to Dr. Herbert Dreeskamp for helpful suggestions and for giving me insights into the brilliance and work ethics of Dr. Theodor F€orster. References 1 (a) F€ orster, T. (1946) Naturwissenschaften, 2 3 4 5 6 7 8 9 10 11 12 13 14 33, 166–175; (b) English translation by Suhling, K. (2012) Journal of Biomedical Optics, 17, 01102-1–01102-10. (a) F€ orster, T. (1948) Annalen der Physik, 2, 55–75; (b) English translation by Mielczarek, E.V., Greenbaum, E., and Knox, R.S. (eds) (1993) Biological Physics, American Institute of Physics, New York, pp. 148–160. F€ orster, T. (1959) Discussions of the Faraday Society, 27, 7–17. €r F€ orster, T. (1949) Zeitschrift fu Naturforschung, 4a, 321–327. F€ orster, T. (1965) in Modern Quantum Chemistry: Istanbul Lectures Part II. Action of Light and Organic Crystals (ed. O. Sinano glu), Academic Press, New York, pp. 93–137. Van der Meer, B.W., Coker, G., III., and Chen, S.-Y.S. (1994) Resonance Energy Transfer: Theory and Data, Wiley-VCH Verlag GmbH, Wienheim. Clegg, R.M. (2006) Reviews in Fluorescence, vol. 3 (eds C.D Geddes and J.R. Lakowicz), Springer Science þ Business Media Inc., pp. 1–45. http://en.wikiquote.org/wiki/Isaac_Newton, Wikipedia, 2002, accessed 2012. Cario, G. (1922) Zeitschrift f€ ur Physik, 10, 185–199 (in German). Cario, G. and Franck, J. (1922) Zeitschrift f€ ur Physik, 11, 161–166 (in German). Franck, J. (1922) Zeitschrift f€ ur Physik, 9, 259–266 (in German). Kallmann, H. and London, F. (1928) Zeitschrift f€ ur physikalische Chemie, B2, 207–243 (in German). Perrin, J. (1925) Conseil de Chimie, Solvay, 2iem, Paris, 1924 Paris, Gauthier & Villar, French, pp. 322–398. Perrin, J. (1927) Comptes Rendus Hebdomadaires des Seances de l’Academie des Sciences, 184, 1097–1100 (in French). 15 Perrin, F. (1932) Annales de Chimie et de 16 17 18 19 20 21 22 23 24 25 26 27 28 Physique (Paris), 17, 283–314 (in French). Perrin, F. (1933) Annales de l’Institut Henri Poincare, 3, 279–318 (in French). Berberan-Santos, M.N. (2001) New Trends in Fluorescence Spectroscopy. Applications to Chemical and Life Sciences (eds B. Valeur and J.-C. Brochon), Springer, Berlin, 7–33. Oppenheimer, J.R. (1941) Physical Review, 60, 158. (a) Arnold, W. and Oppenheimer, J.R. (1950) The Journal of General Physiology, 33, 423–435; (b) See also Knox, R.S. (1995) Photosynthesis Research, 48, 35–39 for a review of the contributions by Arnold. Knox, R.S. (2012) Journal of Biomedical Optics, 17, 01103-1–01103-6. Braslavsky, S.E., Fron, E., Rodríguez, H. B., San Roman, E., Scholes, G.D., Schweitzer, G., Valeur, B., and Wirz, J. (2008) Photochemical & Photobiological Sciences, 7, 1444–1448. Berberan-Santos, M.N. and Prieto, M.J.E. (1987) Journal of the Chemical Society: Faraday Transactions II, 83, 1391–1409. Andrews, D.L. and Rodriguez, J. (2007) Journal of Chemical Physics, 127, 0845091–084509-7. Steinberg, I.Z. (1971) Annual Review of Biochemistry, 40, 83–114. Dexter, D.L. (1953) Journal of Chemical Physics, 21, 836–850. Knox, R.S. and van Amerongen, H. (2002) The Journal of Physical Chemistry B, 106, 5289–5293. Chang, J.C. (1977) Journal of Chemical Physics, 67, 3901–3909. Scholes, G.D. (1999) Resonance Energy Transfer (eds D.L. Andrews and A.A. Demidov), John Wiley & Sons, Ltd, Chichester, pp. 212–243. j59 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 3 F€orster Theory 29 Vanbeek, D.B., Zwier, M.C., Shorb, J.M., 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 and Krueger, B.P. (2007) Biophysical Journal, 92, 4168–4178. nas, G. and Andrews, D.L. (1999) Juzeliu Resonance Energy Transfer (eds D.L. Andrews and A.A. Demidov), John Wiley & Sons, Ltd, Chichester, pp. 65–107. http://scienceworld.wolfram.com/ physics/ElectricMultipoleExpansion.html. Weisstein, E., 2002, accessed 2012. F€ orster, T. (1951) Fluoreszenz Organischer Verbindunger, Vandenhoeck & Ruprecht, G€ ottingen. http://www.tf.uni-kiel.de/matwis/amat/ mw1_ge/kap_2/basics/b2_1_14.html. Foell, H., 1998, accessed 2012. http://www.youtube.com/watch? v¼PlpXrB4NP3w, Burke, G., 2007, accessed 2012. http://www.youtube.com/watch? v¼uWoiMMLIvco, Barrett, B., 2011, accessed 2012. Kasha, M. (1991) Physical and Chemical Mechanisms in Molecular Radiation Biology (eds W.A. Glass and M.N. Varma), Plenum Press, New York, pp. 231–255. Knox, R.S. (1975) Bioenergetics of Photosynthesis (ed. Govindjee), Academic Press, New York, pp. 183–221. Knox, R.S. (1968) Physica, 39, 361–386. Craver, F.W. and Knox, R.S. (1971) Molecular Physics, 22, 385–402. Craver, F.W. (1971) Molecular Physics, 22, 404–420. (a) Bojarski, C. (1972) Journal of Luminescence, 5, 413–429; (b) Bojarski, C. (1974) Journal of Luminescence, 9, 40–45; (c) Bojarski, C. (1979) Journal of Luminescence, 20, 373–378. Bodunov, E.N. (1981) Optics and Spectroscopy, 50, 553–554. Baumann, J. and Fayer, M.D. (1986) Journal of Chemical Physics, 85, 4087– 4107. Gochanour, C.R., Anderson, H.C., and Fayer, M.D. (1979) Journal of Chemical Physics, 70, 4254–4271. Kawski, A. (1983) Photochemistry and Photobiology, 38, 487–508. Bojarski, C. and Sienicki, K. (1989) Photochemistry and Photophysics, vol. 1 (ed. J.F. Rabek), CRC Press, Boca Raton, FL, pp. 1–57. Beddard, G.S. (1983) Time-Resolved Fluorescence Spectroscopy in Biochemistry and 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 Biology (eds R.B. Cundall and R.E. Dale), Plenum Press, New York, pp. 451–461. Eyring, G. and Fayer, M.D. (1984) Journal of Chemical Physics, 81, 4314–4321. Huber, D.L., Hamilton, D.S., and Barnett, B. (1977) Physical Review B, 16, 4624–4650. Berberan-Santos, M.N., Nunes Pereira, E.J., and Martinho, J.M.G. (1997) Journal of Chemical Physics, 107, 10480–10484. Berberan-Santos, M.N., Nunes Pereira, E.J., and Martinho, J.M.G. (1999) Resonance Energy Transfer (eds D.L. Andrews and A.A. Demidov), John Wiley & Sons, Ltd, Chichester, pp. 108–149. Hart, D.E., Anfinrud, P.A., and Struve, W.S. (1987) Journal of Chemical Physics, 86, 2689–2696. Berberan-Santos, M.N. and Valeur, B. (1991) Journal of Chemical Physics, 95, 8048–8055. Valeur, B. and Berberan-Santos, M.N. (2012) Molecular Fluorescence: Principles and Applications, 2nd edn, Wiley-VCH Verlag GmbH, Weinheim. Fung, B.K.-K. and Stryer, L. (1978) Biochemistry, 17, 5241–5248. Wolber, P.K. and Hudson, B.S. (1979) Biophysical Journal, 28, 197–210. Estep, T.N. and Thompson, T.E. (1979) Biophysical Journal, 26, 195–208. Dewey, T.G. and Hammes, G.G. (1980) Biophysical Journal, 32, 1023–1036. Snyder, B. and Freire, E. (1982) Biophysical Journal, 40, 137–148. Doody, M.C., Sklar, L.A., Pownall, H.J., Sparrow, J.T., Gotto, A.M., Jr., and Smith, L.C. (1983) Biophysical Chemistry, 17, 139– 152. Kellerer, H. and Blumen, A. (1984) Biophysical Journal, 46, 1–8. Shaklai, N., Yguerabide, J., and Ranney, H.M. (1977) Biochemistry, 16, 5585–5592. Koppel, D.E., Fleming, P.J., and Strittmatter, P. (1979) Biochemistry, 18, 5450–5457. Yguerabide, M. (1994) Biophysical Journal, 66, 683–693. Runnels, L.W. and Scarlata, S.F. (1995) Biophysical Journal, 69, 1569–1583. Towles, K.B., Brown, A.C., Wrenn, S.P., and Dan, N. (2007) Biophysical Journal, 93, 655–667. Kleinfeld, A.M. (1985) Biochemistry, 24, 1874–1882. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 60 68 Kleinfeld, A.M. and Lukacovic, 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 M.F. (1985) Biochemistry, 24, 1883–1890. Fleming, P.J., Koppel, D.E., Lau, A.L., and Strittmatter, P. (1978) Biochemistry, 18, 5458–5464. Wolf, D.A., Winiski, A.P., Ting, A.E., Bocian, K.M., and Pagano, R.E. (1992) Biochemistry, 31, 2865–2873. Davenport, L., Dale, R.E., Bisby, R.H., and Cundall, R.B. (1985) Biochemistry, 24, 4097–4108. Dobretsov, G.E., Kurek, N.K., Machov, V.N., Syrejshchikova, T.I., and Yakimenko, M.N. (1989) Journal of Biochemical and Biophysical Methods, 19, 259–274. John, E. and J€ahnig, F. (1991) Biophysical Journal, 60, 319–328. Fernandes, F., Neves, P., Gameiro, P., Loura, L.M.S., and Prieto, M. (2007) Biochimica et Biophysica Acta, 1768, 2822–2830. Coutinho, A., Loura, L.M.S., Fedorov, A., and Prieto, M. (2008) Biophysical Journal, 95, 4726–4736. Davenport, L. (1997) Methods in Enzymology, 278, 487–512. Gnanakaran, S., Haran, G., Kumble, R., and Hochstrasser, R.M. (1999) Resonance Energy Transfer (eds D.L. Andrews and A. A. Demidov), John Wiley & Sons, Ltd, Chichester, pp. 308–365. Van Grondelle, R. and Somsen, O.J.G. (1999) Resonance Energy Transfer (eds D. L. Andrews and A.A. Demidov), John Wiley & Sons, Ltd, Chichester, pp. 366–398. Demidov, A.A. (1999) Resonance Energy Transfer (eds D.L. Andrews and A.A. Demidov), John Wiley & Sons, Ltd, Chichester, pp. 435–465. Yang, M., Damjanovic, A., Vaswani, H.M., and Fleming, G.R. (2003) Biophysical Journal, 85, 140–158. Redfield, A.G. (1965) Advances in Magnetic and Optical Resonance, 1, 1–32. Craig, D.P. and Walmsley, S.H. (1968) Excitons in Molecular Crystals, Benjamin, New York. Davydov, A.S. (1948) Journal of Experimental and Theoretical Physics (USSR), 18, 210–218. Kasha, M., Rawls, H.R., and El-Bayoumi, M.A. (1966) Pure and Applied Chemistry, 11, 371–392. 85 McClure, D.S. (1959) Electronic Spectra of 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 Molecules and Ions in Crystals, Academic Press, New York. Buckingham, A.D. and Dalgarno, A. (1952) Proceedings of the Royal Society of London A, 213, 327–349. Merrifield, R.E. (1955) Journal of Chemical Physics, 23, 402–412. Koutecky, J. and Paldus, J. (1962) Collection of Czechoslovak Chemical Communication, 27, 599–617. Koutecky, J. and Paldus, J. (1963) Theoretica Chimica Acta (Berlin), 1, 268–281. Andrews, D.L. (1989) Chemical Physics, 135, 195–201. Avery, J. (1984) International Journal of Quantum Chemistry, 135, 195–201. Craig, D.P. and Thirunamachandran, T. (1984) Molecular Quantum Electrodynamics, Academic Press, New York. McLone, R.R. and Power, E.A. (1964) Mathematika, 11, 91–94. Scholes, G.D. and Andrews, D.L. (1997) Journal of Chemical Physics, 107, 5374– 5384. Azumi, T. and McGlynn, S.P. (1965) Journal of Chemical Physics, 42, 1675– 1680. Murrell, J.N. and Tanaka, J. (1964) Molecular Physics, 7, 363–380. Naqvi, K.R. (1981) The Journal of Physical Chemistry, 85, 2303–2304. Naqvi, K.R. and Steel, C. (1970) Chemical Physics Letters, 6, 29–32. Harcourt, R.D., Scholes, G.D., and Ghiggino, K.P. (1994) Journal of Chemical Physics, 101, 10521–10525. Scholes, G.D. and Ghiggino, K.P. (1994) The Journal of Physical Chemistry, 98, 4580–4590. Scholes, G.D. and Harcourt, R.D. (1996) Journal of Chemical Physics, 104, 5054–5061. Scholes, G.D., Harcourt, R.D., and Ghiggino, K.P. (1995) Journal of Chemical Physics, 102, 9574–9581. Scholes, G.D., Harcourt, R.D., and Fleming, G.R. (1997) The Journal of Physical Chemistry, 101, 7302–7312. Andrews, D.L., Bradshaw, D.S., Jenkins, R.D., and Rodriguez, J. (2009) Dalton Transactions, 45, 10006–10014. Huo, P. and Coker, D.F. (2010) Journal of Chemical Physics, 133, 184108. j61 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 3 F€orster Theory 106 Hauser, M., Klein, U.K.A., and G€ osele, 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 €r physikalische U. (1976) Zeitschrift fu Chemie, 101, 255–266. Blumen, A., Klafter, J., and Zumofen, G. (1986) The Journal of Physical Chemistry, 84, 1397–1401. Duportail, G., Merola, F., and Cianos, P. (1995) Journal of Photochemistry and Photobiology A: Chemistry, 89, 135–140. Tcherkasskaya, O., Gronenborn, A.M., and Klushin, L. (2002) Biophysical Journal, 83, 2826–2834. Drake, J.M., Klafter, J., and Levitz, P. (1991) Science, 251, 1574–1579. Dewey, T.G. (1992) Accounts of Chemical Research, 85, 4087–4107. Haas, E., Katchalski-Katzir, E., and Steinberg, I.Z. (1978) Biopolymers, 17, 11–31. Thomas, D.D., Carlsen, W.F., and Stryer, L. (1978) Proceedings of the National Academy of Sciences of the United States of America, 75, 5746–5750. Katchalski-Katzir, E. and Steiberg, I.Z. (1981) Advances in Optical Biopsy and Optical Mammography, 366, 41–61. Steinberg, I.Z., Haas, E., and KatchalskiKatzir, E. (1983) in Time-Resolved Fluorescence Spectroscopy in Biochemistry and Biology (eds R.B. Cundall and R.E. Dale), Plenum Press, New York, pp. 411–450. Fairclough, R.H. (2006) Biophysical Journal, 91, 1143–1144. Stryer, L., Thomas, D.D., and Meares, C.F. (1982) Annual Review of Biophysics and Bioengineering, 11, 203–222. Kouyama, T.K., Kinosita, K., Jr., and Ikegami, A. (1983) Journal of Molecular Biology, 165, 91–107. Mersol, J.V., Wang, H., Gafni, A., and Steel, D.G. (1992) Biophysical Journal, 61, 1647–1655. G€ osele, U., Hauser, M., Klein, U.K.A., and Frey, R. (1975) Chemical Physics Letters, 24, 519–523. Beechem, J.M. and Haas, E. (1989) Biophysical Journal, 55, 1225–1236. Lakowicz, J.R., Szmacinski, H., Gryczynski, I., Wick, W., and Johnson, M. 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 L. (1990) The Journal of Physical Chemistry, 94, 8413–8416. Kusba, J., Li, L., Gryczynski, I., Piszczek, G., Johnson, M., and Lakowicz, J.R. (2002) Biophysical Journal, 82, 1358–1372. Haas, E. and Steinberg, I.Z. (1984) Biophysical Journal, 46, 429–437. Tanaka, F. and Mataga, N. (1975) Biophysical Journal, 39, 129–140. Tanaka, F. and Mataga, N. (1975) Photochemistry and Photobiology, 29, 1091–1097. Szabo, A. (1984) Journal of Chemical Physics, 81, 150–167. Cross, A.J., Waldeck, D.H., and Fleming, G.R. (1983) Journal of Chemical Physics, 78, 6455–6467. Van der Meer, B.W., Raymer, M.A., Wagoner, S.L., Hackney, R.L., Beechem, J.M., and Gratton, E. (1993) Biophysical Journal, 64, 1243–1263. Canley, B.A., Brown, F.L.H., and Lipman, E.A. (2009) Journal of Chemical Physics, 131, 104509. Raicu, V. (2007) Journal of Biological Physics, 33, 109–127. Bojarski, P., Kulak, L., Walczewska-Szewc, K., Synak, A., Marzullo, V.M., Luini, A., and D’Auria, S. (2011) The Journal of Physical Chemistry B, 115, 10120–10125. Rolinski, O.J. and Birch, D.J.S. (2000) Journal of Chemical Physics, 112, 8923–8933. Rolinski, O.J. and Birch, D.J.S. (2008) Journal of Chemical Physics, 129, 144507. Swathi, R.S. and Sebastian, K.L. (2010) Journal of Chemical Physics, 132, 104502. Wong, C.Y., Curutchet, C., Tretiak, S., and Scholes, G.D. (2009) Journal of Chemical Physics, 130, 081104. Geddes, C.D. (ed.) (2010) Metal-Enhanced Fluorescence, John Wiley & Sons, Inc., New York. Lakowicz, J.R. (2005) Analytical Biochemistry, 337, 171–194. Klan, P. and Wirz, J. (2009) Photochemistry of Organic Compounds: From Concepts to Practice, John Wiley & Sons, New York. http://www.wiley.com/WileyCDA/ WileyTitle/productCd-1405161736.html Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 62 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements B. Wieb van der Meer, Daniel M. van der Meer, and Steven S. Vogel 4.1 Two-Thirds or Not Two-Thirds? Two-thirds or not two-thirds? This is the question. Kappa-squared can vary between 0 and 4, but when the orientations of donor and acceptor dipoles randomize within the lifetime of the excited state, its value is 2/3. Many authors of FRET papers adopt this assumption. However, there is strong evidence that kappa-squared is definitely not equal to 2/3 in many cases. For example, in Ref. [1], a series of DNA conjugates in which a donor (stilbene dicarboxamide) and an acceptor (perylene dicarboxamide) are covalently attached to opposite sites of an A:T base pair duplex domain consisting of 4–12 base pairs yield a FRET efficiency that is strongly nonlinear with varying distance. For 7–9 base pairs the efficiency drops to almost zero consistent with a near-zero value of kappa-squared; whereas for 5 and 10 base pairs the efficiency reaches a maximum consistent with a kappa-squared value of 1 [1]. In another example [2], a Cy3 donor and a Cy5 acceptor are attached to the 50 -termini of duplex DNA via a 3-carbon linker to the 50 -phosphate so that they are predominantly stacked onto the ends of the helix in the manner of an additional base pair [2]. A cartoon illustrating the first two examples is shown in Figure 4.1a; the third example is shown in Figure 4.1b. The transition dipoles are essentially perpendicular to the helical axis, and the periodicity is on the order of 5 base pairs. As a result, kappasquared changes dramatically with the donor–acceptor distance, approaching zero at 13 and 18 base pairs. The graph of FRET efficiency versus donor–acceptor distance looks like the graph of the height of a bouncing ball versus time (dashed curve in Figure 4.18). In reality, there is some motional averaging so that for none of the base pair choices does the efficiency dip to zero, but there are clear maxima and minima in the efficiency versus distance curve at predictable donor–acceptor distances (we will come back to this trend with Figure 4.18 in Section 4.9). The error in the distance by assuming kappa-squared ¼ 2/3 is about 25% at 13 base pairs [2]. In a third example [3], a Cy3 donor and a Cy5 acceptor are rigidly attached to DNA in such a way that the dipoles are essentially parallel to the axis of the DNA molecule. In two cases, a configuration of collinear donor and acceptor dipoles was engineered: one with Cy3 and Cy5 on the same B-DNA strand and separated by three helical turns FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j63 j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Figure 4.1 (a) A cartoon for the FRET situation in Refs [1,2]. The double-headed arrows represent transition moments for donor or acceptor. The angle between them depends on the donor–acceptor distance relative to the helical pitch, reaching near-zero (k2 1) when the distance equals an even integer times a quarter pitch and 90 when the distance equals an odd integer times a quarter pitch (k2 0). (b) A cartoon for the FRET situation in Ref. [3] (samples 1 and 2) where the transition moments are essentially aligned with the donor–acceptor separation vector corresponding to k2 4. (sample 1 in Ref. [3]) and the other with Cy3 and Cy5 on opposite strands and separated by 2.5 turns (sample 2 in Ref. [3]). In both cases, the two oscillating dipoles are expected to be collinear for which configuration kappa-squared reaches its maximum value of 4. The experimentally obtained kappa-squared values (from the known distance, the measured efficiency, and the known F€ orster distance R0 ¼ 6:34ðk2 Þ1=6 nm) was 3.2 for sample 1 and 3.5 for sample 2, indicating the near-parallel alignment of the dipoles with the line connecting donor and acceptor [3]. These three examples of having a kappa-squared different from 2/3 are for traditional donors and acceptors attached to DNA. Furthermore, the 2/3 assumption also fails in FRET experiments using fluorescent proteins as donors and acceptors, which undergo, usually restricted, rotation independent of each other that is slow relative to the lifetime of the excited state in the presence of acceptor. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 64 4.2 Relevant Questions Kappa-squared needs attention, but this orientation factor problem does not need to be disorienting! A few relevant questions must be asked. Do the orientations of donor and acceptor change during the time within which transfer may take place, that is, effectively during the lifetime of the excited state in the presence of the acceptor(s)? And, if they do, how? Do they change rapidly during this time? Does the dynamic averaging regime apply or the static averaging regime or neither? Is the fluorescence polarized? To what extent? Can kappa-squared be measured? Can depolarization factors be measured? Are simulations available or is structural information obtainable that may exclude certain orientations? First, we need to know how to visualize kappa-squared. 4.3 How to Visualize Kappa-Squared? Kappa-squared for a given donor–acceptor pair depends on the direction of the emission transition moment of the donor, the absorption transition moment of the acceptor, and the line connecting the centers of the donor and the acceptor. We can introduce unit vectors: ^d along the emission transition moment of the donor, ^ a along the absorption transition moment of the acceptor, and ^r pointing from the center of the donor to the center of the acceptor. These three unit vectors are shown in Figure 4.2. To visualize better the implications for kappa-squared of this three-dimensional geometry, the following exercise is recommended. Hold your two index fingers in front of your face and simulate the donor dipole with your left index finger and that ^ a^, and ^r ; d^ is Figure 4.2 The unit vectors d, along the emission transition moment of the donor, ^a is along the absorption transition moment of the acceptor, and ^r points from the center of the donor to the center of the acceptor. The d^ and ^a vectors are displayed at arbitrary orientations within the spheres represented at their locations. j65 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.3 How to Visualize Kappa-Squared? j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements of the acceptor with the other index finger. You can then rotate one hand in three dimensions around its wrist and estimate (see below) how kappa-squared changes for the given orientation of the other finger and for any of the orientations of the finger of the one hand, and then rotate the other hand while maintaining the absolute orientations of the fingers constant and evaluate the orientation factor again. Rotating the hands around each other demonstrates that significant changes in kappa-squared also result for transfer in different directions, that is, for different orientations of the separation vector ^r. The angle between ^ d and ^ a is qT , that between ^d and ^r is qD , and that between ^a and ^r is qA . There are three common ways of expressing kappa-squared (k2 ) in angles: k2 ¼ ðcos qT 3 cos qD cos qA Þ2 ; ð4:1Þ 2 k2 ¼ ðsin qD sin qA cos w 2 cos qD cos qA Þ ; ð4:2Þ k ¼ ð1 þ cos qD Þcos v; ð4:3Þ 2 2 2 where w is the angle between the projections of ^ d and ^ a on a plane perpendicular to ^r and v is the angle between the electric dipole field due to the donor at the location of d, and the unit vector along the acceptor and ^a. The electric field is along 3^r cos qD ^ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi this direction is ^eD ¼ ð3^r cos qD ^dÞ= 1 þ 3 cos2 qD . The angles appearing in Equations 4.1–4.3 are illustrated in Figure 4.3. The dependence of kappa-squared on cos qD and cos v expressed in Equation 4.3 is illustrated in Figure 4.4. Figure 4.3 illustrates in particular the different planes formed by the vectors: the DR plane through ^d and ^r , the AR plane through ^ a and ^r , the DA plane through ^ d and ^a, and the EDA plane through ^eD and ^a. Note that ^eD lies always in the DR plane and that, whenever ^a is perpendicular to this plane, ^ a is also perpendicular to ^ d, ^r , and ^eD , so that kappa-squared is zero. Equation 4.3 gives insight into the distribution of kappa-squared values. The highest value possible, 4, can only be realized if ^r and ^ d Figure 4.3 In this illustration of the angles in Equations 4.1–4.3, uT ¼ 97:18 , uD ¼ uA ¼ 60 , ^ ^r , and ^eD are in the DR w ¼ 120 , and v ¼ 48:59 , yielding k2 ¼ 0:7625. The unit vectors d, plane, ^a and ^r in the AR plane, and, ^eD and ^a are in the EDA plane. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 66 Figure 4.4 Kappa-squared versus the absolute values of cos uD and cos v. are parallel or antiparallel yielding cos2 qD ¼ 1 and if, at the same time, ^ a and ^eD are parallel or antiparallel, ensuring that cos2 v is also equal to 1. On the other hand, whenever ^a and ^eD are perpendicular to each other, kappa-squared equals zero. Therefore, if we consider all possible orientations, there is a very high probability that kappa-squared is low and a much smaller probability that it is high. Accordingly, the isotropic average is 2/3: if all orientations are equally probable in three dimensions, the average of cos2 v is 1/3 and that of cos2 qD is also equal to 1/3, so Equation 4.3 predicts the isotropically averaged value of kappa-squared as 2/3. Expressions of kappa-squared in terms of unit vectors and dot products are also relevant: k2 ¼ ð^a ^d 3ð^a ^r Þð^r ^dÞÞ2 ¼ ð^a ^eD Þ2 ð1 þ 3ð^r ^ dÞ2 Þ2 : ð4:4Þ The right-hand side is the vector form of Equation 4.3 and the expression in the middle is the vector form of Equation 4.1. From these forms, it is clear that kappa-squared does not change if we 1) 2) 3) 4) flip the donor transition moment, ^d ! ^d, flip the acceptor transition moment, ^a ! ^a, allow the donor and acceptor to trade places, ^r ! ^r , and interchange the donor and acceptor transition moments, ^ a$^ d. The transition moments can be visualized as rod-like molecular antennas or even index fingers. It is instructive to choose different orientations and evaluate the corresponding kappa-squared values as in the examples shown in Figure 4.5. j67 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.3 How to Visualize Kappa-Squared? j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Figure 4.5 Examples of donor and acceptor orientations with corresponding k2 values. The donor dipole is along the bar in the center of each circle with various examples of acceptor dipoles, also depicted as bars, along the circumference. For the first circle (a), the donor and acceptor transition moments are parallel to each other in the same plane and k2 varies between 4 and 0, depending on the location of the acceptor on the circle. These values are labeled next to the acceptor. Note that even when the donor and acceptor transition moments are parallel, k2 can still be zero. For the circle (b), with donor and acceptor dipoles again lying in the same plane, the acceptor dipole is oriented along the electric field of the excited donor at the location of the acceptor with k2 values between 1 and 4. For the circles (c) and (d), the orientation factor is 0 for each example as the acceptor dipole is perpendicular to the donor electric field. For the circle (c), the donor and the acceptor are oriented in the same (DR) plane, but for the circle (d), the acceptor is perpendicular to the DR plane. In each circle, the electric field lines of the excited donor are shown. 4.4 Kappa-Squared Can Be Measured in At Least One Case Dale has shown that for the special case depicted in Figure 4.6, kappa-squared can be measured using time-resolved fluorescence depolarization [4]. The donor and acceptor are assumed not to move with respect to the macromolecule, but the whole system can rotate around its axes, exhibiting rotational diffusion around the symmetry axis at D== and around any axis perpendicular to that at D? , where D== and D? are rotational diffusion constant. If both donor and acceptor fluoresce, three different time-resolved anisotropies can be measured: r D ,r A , and r T . r D is the donor fluorescence anisotropy (donor is excited and donor fluorescence is measured), r A is the acceptor fluorescence anisotropy (acceptor is excited and acceptor fluorescence is measured), and r T is the transfer anisotropy (donor is excited and sensitized acceptor fluorescence is observed). These anisotropies vary with t, the time after an ultrashort flash, and are given by the following: r D ¼ r D ðtÞ ¼ b1D eð2D? þ4D== Þt þ b2D eð5D? þD== Þt þ b3D e6D? t : b1D ¼ 3 3 sin4 qD b2D ¼ sin2 2 qD ; 10 10 b3D 4 3 2 1 2 ¼ : cos qD 10 2 2 ð4:5aÞ ð4:5bÞ Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 68 Figure 4.6 A cartoon of a macromolecule with a donor (D, emission transition dipole along ^ and an acceptor (A, the unit vector d) absorption transition dipole along the unit vector ^a) rigidly attached to the symmetry axis of the macromolecule, which undergoes rotational diffusion around the symmetry axis at D== and around any axis perpendicular to that at D? . D== and D? are rotational diffusion constants. The angles appearing in the anisotropy decays (uD , uA , uT , and w) are as shown in Figure 4.4 and defined near Equation 4.1. This cartoon is similar to the one presented by Dale [4]. r A ¼ r A ðtÞ ¼ b1A eð2D? þ4D== Þt þ b2A eð5D? þD== Þt þ b3A e6D? t : b1A ¼ 3 sin4 qA ; 10 b2A ¼ 3 sin2 2qA ; 10 b3A 4 3 2 1 2 ¼ : cos qA 10 2 2 r T ¼ r T ðtÞ ¼ b1T eð2D? þ4D== Þt þ b2T eð5D? þD== Þt þ b3T e6D? t : 3 sin2 qD sin2 qA cos 2w 10 h i 3 ¼ 2ðcos qT cos qD cos qA Þ2 sin2 qD sin2 qA : 10 ð4:5cÞ ð4:5dÞ ð4:5eÞ b1T ¼ ð4:5f ðiÞÞ 3 6 sin 2qD sin 2qA cos w ¼ cos qD cos qA ½cos qT cos qD cos qA : 10 5 ð4:5f ðiiÞÞ 4 3 2 1 3 2 1 ¼ cos qD cos qA : ð4:5f ðiiiÞÞ 10 2 2 2 2 b2T ¼ b3D Global analysis allows one to obtain the rotational diffusion constants and cos qD (from r D ), cos qA (from r A ), and cos qT (from r T ), so that kappa-squared can be calculated using Equation 4.1 [4]. The conclusion for now is that kappa-squared can j69 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.4 Kappa-Squared Can Be Measured in At Least One Case j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements be measured in this specific case. However, this approach can probably be extended to more and perhaps all cases. This particular case is unique in that the orientations of the transition moments are fixed in the frame of the macromolecule. In most cases, some motion occurs. The range and frequencies of such motions may differ. The concept of averaging regimes is useful for understanding the implications of motion for FRET. 4.5 Averaging Regimes The initial steps of a FRET experiment involve the absorption of a photon by a donor fluorophore. Absorption of a photon is rapid, typically occurring within a femtosecond (1015 s), and results in the elevation of a ground-state electron into a myriad of potential electronic and vibrational excited states. Over the next few hundred femtoseconds, this array of potential excited-state electronic and rotational–vibrational energy sublevels are consolidated into the Boltzmann rotational–vibrational level manifold of the lowest-energy singlet excited state, as a result of vibrational energy loss due to subsequent kinetic interactions between the excited fluorophore and surrounding molecules. Fluorophores, in general, spend from picoseconds to tens of nanoseconds in this relatively long-lived lowest singlet excited state before eventually transitioning back to a ground-state sublevel. With their return to a ground state, excess excited-state energy will be either emitted as a photon (donor fluorescence), transferred to a nearby acceptor (FRET), or it will be utilized by some other nonradiative mechanisms. To understand the factors that can influence the probability of energy transfer by FRET, one must understand the types of events that can occur while a fluorophore is in its excited state. In relation to kappa squared, the main factors that must be considered is to what extent donor and acceptor fluorophores can move relative to each other while in the excited state – specifically, how fluorophore motion may influence the position of an acceptor relative to the orientation of the donor emission dipole, and how it may influence the orientation of the acceptor absorption dipole relative to the orientation of the donors excited-state electric field. When every donor and every acceptor can take up its entire range of orientations during the lifetime of the excited donor state in the presence of acceptors, the system is said to be in the dynamic averaging regime, and the dynamic averaging condition applies. In this regime, kappa-squared can be replaced by an appropriate average value, and the average FRET efficiency is given by ð3=2Þhk2 iR 0 ; 6 2 ð3=2Þhk iR0 þ r 6 6 hE idynamic ¼ ð4:6Þ DA 0 is the F€ where the brackets denote an average, R orster distance when k2 ¼ 2=3, and r DA is the donor–acceptor distance. The isotropic condition applies when all orientations are equally probable. The dynamic isotropic average of kappa-squared equals 2/3 in the one-, two-, and three-dimensional cases [5]. As discussed above, the Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 70 isotropic average is not always valid. The dynamic averaging regime is discussed further in Sections 4.6–4.9. When the rates of rotation are small compared to the rate of donor decay in the presence of acceptor, the system is in the static averaging regime, and the static averaging condition applies. In this regime, kappa-squared cannot be replaced by a universal average value, and the average FRET efficiency is given by * hE istatic ¼ ð3=2Þk2 R 0 6 0 þ r6 ð3=2Þk2 R + 6 ; ð4:7Þ DA where the symbols have the same meaning as in Equation 4.6. Note, however, the difference between calculating an average static efficiency and obtaining an average 60 Þ, the dynamic efficiency. Also note that at large distances when r 6DA ð3=2Þðk2 R differences between Equations 4.7 and 4.6 vanish. Effective values for kappasquared in the static and dynamic averaging regimes have been derived by Dale for random spatial distributions of separations of free donors and acceptors in solutions of three, two, and one dimension with, as appropriate, random three- and two-dimensional orientational distributions or, for the one-dimension spatial, onedimension orientational case, the inline configuration [4,6]: Orientational distribution 3D Spatial solution distribution 3D Dynamic average of kappa-squared 2/3 2D 3D 2/3 1D 3D 2/3 2D 2D 5/4 1D 2D 5/4 1D 1D 4 Static average of kappa-squared DpffiffiffiffiffiE2 k2 ffi 0:69012 ffi 0:4762 Dpffiffiffiffiffi 3 3 k2 i ffi 0:73973 ffi 0:4048 Dpffiffiffiffiffi 6 6 k2 i ffi 0:83056 ffi 0:3281 Dpffiffiffiffiffi 3 3 k2 i ffi 0:94622 ffi 0:8471 Dpffiffiffiffiffi 6 6 k2 i ffi 0:94566 ffi 0:7151 Dpffiffiffiffiffi 6 6 k2 i ¼ 4 In the inline configuration (1D, 1D), a single value for kappa-squared applies, so that in this case the dynamic and static values are identical. The first report on FRET in 2D free donor, free acceptor solutions was published by Tweet et al. [7]. Loura et al. [8] confirmed 1D solution FRET in an experimental system, with the theory given in detail in Ref. [9]. It is possible that the average rate of transfer is on the same order of magnitude as a dominant rate of rotation for the donor or acceptor. In this case, the system is neither in the dynamic regime nor in the static regime. This case is discussed in Section 4.12. j71 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.5 Averaging Regimes j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements 4.6 Dynamic Averaging Regime Dale et al. [10] and Haas et al. [11] have shown that in the dynamic averaging regime, fluorescence depolarization data allow one to remove some of the uncertainty in the FRET distance resulting from kappa-squared. Dale et al. emphasize depolarization because of rapid restricted rotations [10], whereas Haas et al. [11] mainly consider excitation into overlapping transitions as the reason for low polarization values. However, in the dynamic averaging regime, depolarization due to degeneracy or overlap of transitions is essentially indistinguishable from depolarization resulting from reorientations. Indeed, it has been shown that the equation for the average kappasquared derived by Dale et al., Equation 4.10, is the same as the one derived by Haas et al., if cylindrical symmetry in the transitions is assumed [12]. In the dynamic averaging regime, the donor emission moment, which is along the unit vector ^ d, fluctuates rapidly X around ^d (unit vector, called donor axis) and the acceptor absorption moment, which is along the unit vector ^a, fluctuates rapidly around ^ aX (unit vector, called acceptor axis). These fluctuations may represent rapid restricted rotations or coupling between overlapping transitions. As a result, the kappa-squared value fluctuates around an average value. And, this average value, which may be used instead of the kappa-squared value appearing in the FRET efficiency, depends on two parameters d and a (defined below) and on three variables W, HD , and HA (defined below). The parameter d is the axial depolarization factor for the donor emission moment: d ¼ dXD ¼ ðp 3 2 1 3 2 1 cos yD ¼ cos yD sin yD F D ðyD ÞdyD ; 2 2 2 2 ð4:8Þ 0 X where yD is the fluctuating angle between ^ d and ^ d , and F D ðyD Þ is a distribution function. Similarly, the axial depolarization factor for the acceptor absorption moment is ðp X 3 2 1 3 2 1 a ¼ dA ¼ ð4:9Þ cos yA ¼ cos yA sin yA F A ðyA ÞdyA ; 2 2 2 2 0 where yA is the fluctuating angle between ^ a and ^ ax , and F A ðyA Þ is the distribution function. Note that it is assumed here that the distributions are cylindrically symmetrical. The parameters d and a are second rank orientational order parameters with values between 0.5 and 1 : 1 when the transition moment is completely aligned with its axis, 0 when the angle between the transition moment and its axis is equal to the magic angle at all times or when the transition moment is completely random, and 0.5 when the angle between the transition moment and its axis is 90 , that is, when the transition moment is degenerate in a plane perpendicular to the axis. HD is the angle between the donor axis and the line connecting the centers of donor and acceptor, HA is the angle between this connection line and the acceptor axis, and W is the angle between the projections of the donor and acceptor axes on a plane perpendicular to the connection line. The average value of the orientation factor Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 72 in the dynamic averaging regime is [10]: 2 2 1 k ¼ ðd þ aÞ þ dð1 aÞcos2 HD þ að1 dÞcos2 HA þ adk21;1 ; 3 3 with k21;1 ¼ ðsin HD sin HA cos W 2 cos HD cos HA Þ2 : ð4:10aÞ ð4:10bÞ The depolarization factors d and a can be obtained from fluorescence depolarization measurements [10]. The “time-zero” value of the donor fluorescence anisotropy value is proportional to d2 and that of the acceptor is proportional to a2 . Here “time-zero” must be understood in the context of the dynamic averaging regime, and the so-called B€ urkli–Cherry [13] plot illustrates the concept of “zero-time anisotropy.” Figure 4.7 may suggest that time-resolved fluorescence anisotropy measurements are necessary in order to obtain depolarization factors. However, Corry et al. have shown that steady-state confocal microscopy also enables one to measure such factors and that kappa-squared can even be obtained if some knowledge of the relative geometry is assumed [14]. Because of the proportionality between measured fluorescence anisotropy values and the square of d or a, the experimentally obtained depolarization factors can be either positive or negative if d2 and a2 are between 0 and 0.25. These sign ambiguities may be resolved if independent structural or spectroscopic information is available. Figure 4.7 A log–log plot of fluorescence anisotropy versus time after a flash excitation. This is also called a B€ urkli–Cherry plot [13]. It shows a stepwise decrease of the anisotropy with time and nicely illustrates that “zero-time anisotropy” in the context of FRET refers to the anisotropy value reached after the completion of rotations with frequencies higher than the average transfer rate. Note that the lifetime of the excited state (of donor or acceptor) seems not to matter in this graph, but it is relevant in practice because of noise: after a few lifetimes, the time-resolved anisotropy becomes very noisy and is completely unreliable for times larger than about five lifetimes. j73 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.6 Dynamic Averaging Regime j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Dale et al. found the maxima and minima of hk2 i in Equation 4.10 numerically and presented a contour plot allowing to read the highest and lowest hk2 i value for each combination of depolarization factors [10]. It is also possible to find these maxima and minima analytically by setting the derivatives of hk2 i with respect to HD , HA , and W equal to zero, solving the set of three resulting equations [15]. As shown at http:// www.FRETresearch.org, there are six candidates for maxima and minima: k2A ¼ 2 2 2 þ a þ d þ 2ad: 3 3 3 ð4:11aÞ k2P ¼ 2 1 1 a d: 3 3 3 ð4:11bÞ k2H ¼ 2 1 1 a d þ ad: 3 3 3 ð4:11cÞ k2M ¼ 2 1 1 1 þ a þ d ad þ ja dj: 3 6 6 2 ð4:11dÞ k2L ¼ 2 1 1 1 þ a þ d ad ja dj 3 6 6 2 ð4:11eÞ 1 4 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð1 aÞð1 dÞð1 þ 2aÞð1 þ 2dÞ: k2T ¼ ð1 aÞð1 dÞ þ 9 9 ð4:11f Þ The distributions of transition moments can be visualized as ellipsoids with the symmetry axis equal to 1 þ 2d for the donor and 1 þ 2a for the acceptor and with any axis perpendicular to the symmetry axis equal to 1 d for the donor and 1 a for the acceptor. As a result, in the extreme situation where the depolarization factor equals 1, the distribution behaves like a needle-like “molecular antenna,” and in the other extreme where it is 0.5, the transition dipole distribution resembles a disklike “antenna.” Figure 4.8 illustrates the meanings of the six candidates using such ellipsoids and verbal descriptions. Careful comparison (http://www.FRETresearch.org) of the magnitudes of one candidate relative to those of the others in all points of the plane formed by parameter values ð1=2Þ d 1 and ð1=2Þ a 1 leads to the conclusion that there are nine different regions where the maxima and minima can be calculated using the expressions for the six candidates in Equations 4.11a–4.11f. These regions have borders expressed as d ¼ 0, a ¼ 0, C ¼ 0, E ¼ 0, F ¼ 0, or G ¼ 0, where C, E, F, and G are defined as follows: 1 C ¼aþd : 2 ð4:12aÞ E ¼ 3a þ 3d þ 5ad þ 1: ð4:12bÞ F ¼ 2d 3a þ 2ad 1: ð4:12cÞ G ¼ 2a 3d þ 2ad þ 1: ð4:12dÞ The regions are shown in Figure 4.9. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 74 Figure 4.8 Description of the candidates for maxima and minima of the average kappa-squared in the dynamic regime, specified by Equation 4.11. They are also candidates for the most probable kappa-squared in this averaging regime. The meaning of the symbols and the properties of the different regions are specified in Table 4.1. Note that if both depolarization factors, for the donor and the acceptor, are positive, the minimum hk2 i is k2P and the maximum is k2A , as pointed out by Dale A et al. [10]. The reader may wonder why we split up this region in a central 1 P A zone and three sections of 2 around it. The reason is that the six candidates are P not only possible maxima and minima but are also potential answers to the question: j75 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.6 Dynamic Averaging Regime j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Figure 4.9 Regions in the ða; dÞ plane showing in each a column with the kappa-squared maximum indicated by the top letter and the minimum by the bottom letter. Table 4.1 gives details. What is the most probable kappa-squared value? Figure 4.9 also serves as the starting point in our approach to this question. 4.7 What Is the Most Probable Value for Kappa-Squared in the Dynamic Regime? What is the “most probable” kappa-squared value? This is an ambiguous question! If we are trying to find the most probable value per se of kappa-squared, that is, independent of and in isolation from any other FRET parameter, we will get one answer; but if we want the most probable k2 corresponding to the most probable separation derived for a given efficiency, a completely different answer emerges. It is well established that the probability density of kappa-squared per se for a pair of linear donor and acceptor transition moments (a ¼ d ¼ 1 in Equation 4.10a) exhibits an infinitely high peak at k2 ¼ 0 (see Equation 4.21 and Figure 4.19) (also refer to Refs [10,12,16]). Nevertheless, if we consider any nonzero efficiency, however small, whether or not obtained in an actual experimental situation, deriving either from a transfer efficiency or a transfer rate, the most probable kappa-squared Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 76 Table 4.1 Maxima and minima in the dynamic averaging regime. Name of region Maximum hk2 i in that region Minimum hk2 i in that region Definition of region A 1 P k2A k2P fa > 0; d > 0; F G C > 0g Yes A 2 P k2A k2P fa 0; d G C All but not k2T M 3 H k2M k2H fa d M 4 T k2M k2T fa d < 0; E C < 0; F M 5 L k2M k2L fa d < 0; F G H 6 A k2H k2A fa 0; d M 7 A k2M k2A fa d < 0; E H 8 L k2H k2L fa 0; d 0; F G 0g All but not k2T H 9 T k2H k2T fa 0; d 0; F G E < 0g Yes 0; F 0; C Are all candidates valid there? 0g All but not k2T 0g G < 0g All but not k2T 0g 0; E Yes All but not k2T 0g All but not k2T 0g value cannot be equal to zero, because k2 ¼ 0 means that the efficiency also equals zero. Interestingly, the explanation of this paradox is based on the link between kappa-squared and the “relative distance,” which is defined by r “Relative distance” Actual FRET distance ¼ Distance assuming k2 ¼ 2=3 1=6 3 2 : k 2 ð4:13Þ In formal mathematical terms, we need three probability functions for this explanation: the range probability, PR k2min ! k2 , the probability density for kappasquared, pðk2 Þ, and the probability density for the relative distance, Q ðrÞ: P R k2min ! k2 ¼ probability that kappa-squared has a value between k2min and k2 : ð4:14Þ p k2 dk2 ¼ probability that kappa-squared has a value between k2 and k2 þ dk2 : ð4:15Þ Q ðrÞdr ¼ probability that the relative distance has a value between r and r þ dr: ð4:16Þ j77 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.7 What Is the Most Probable Value for Kappa-Squared in the Dynamic Regime? j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Both pðk2 Þ and Q ðrÞ are proportional to the derivative of P R k2min ! k2 . And, because of the link between the relative distance and kappa-squared as expressed in Equation 4.13, it follows that Q ðrÞ ¼ 4r5 p k2 : ð4:17Þ Therefore, the mathematical explanation of the above-mentioned apparent paradox is that the derivative of the pðk2 Þ will in general not be zero when the derivative of Q ðrÞ equals zero because of the factor 4r5 in Equation 4.17. It is difficult to visualize this apparent paradox for the general case of any choice for a and d. However, for the case that both of these depolarization factors are equal to 1, PR k2min ! k2 ¼ PR ð0 ! k2 Þ is the area under the curve that is obtained when cutting the threedimensional plot in Figure 4.3 at a certain kappa-squared level and projecting the cut in the jcos qD j jcos vj plane, as shown in Figure 4.10. In this special case, constant-k2 curves can be calculated from Equation 4.3. The area of the square formed by all possible values of jcos qD j and jcos vj between 0 and 1 represents the total probability of 1 that k2 has any value between 0 and 4, 0 for cos v ¼ 0 and 4 for jcos vj ¼ jcos qD j ¼ 1. We can divide up the square in, say, a hundred pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi strips by drawing the 101 curves jcos vj ¼ k2 =ð1 þ cos 2 qD Þ inside the square choosing k2 equal to 0 4/100, 1 4/100, . . . , 99 4/100, and 100 4/100. This way the area of each strip represents the probability that kappa-squared has a value Figure 4.10 Lines of constant kappa-squared in the jcos uD j jcos vj are shown for the case in which both depolarization factors are equal to 1. The area below the curve labeled 1/3 is equal to the probability that kappa-squared is between 0 and 1/3, the area between this curve and the 2/3 curve represents the probability that kappasquared is between 1/3 and 3/2, and so on. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 78 Figure 4.11 Pair of diagrams illustrating that the question “What is the most probable kappasquared value?” is ambiguous. This example refers to the dynamic regime and a ¼ d ¼ 1. (a) The diagram is for unknown efficiency with k2 ¼ 0 as the most probable value. (b) The diagram corresponds to having a known efficiency and thus a link between the orientation factor and relative distance resulting in k2 ¼ 1 being the most probable kappa-squared value. between the k2 for the lower boundary and that for the upper boundary. Such a division has been initiated in Figure 4.11a. The very first strip between the 0 and 0.04 curves has by far the largest area and successive strips rapidly decrease in area, indicating that the most probable value for the orientation factor is 0. However, nothing is said about the distance or the efficiency. Over that very first strip, the relative distance is 0 at the lower boundary, but 0.626 at the higher boundary, whereas the maximum relative distance is 1.35. As a result, the very first strip in k2 represents 46% of all distance choices. It seems more appropriate, therefore, to translate the combination of a measured efficiency and an independently obtained F€orster distance to a distance with an orientational uncertainty specified by Equation 4.10. In terms of the example of Figure 4.7, this means we should divide up the square by drawing 101 curves jcos vj ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ðð2=3Þr6 Þ=ð1 þ cos 2 qD Þ inside the square by choosing r equal to 0 61=6 =100, 1 61=6 =100, . . . , 99 61=6 =100, and 100 61=6 =100. This way the area of each strip represents the probability that the relative distance has a value between the r for the lower boundary and that for the upper boundary. This is indicated in Figure 4.11b, where careful analysis shows that the strip straddling the left upper corner of the square has the biggest area, corresponding to r ¼ ð3=2Þ1=6 ¼ 1:07 and to k2 ¼ 1. In general, the location of the maximum of Q will differ dramatically from that of the maximum of p, as in the example of Figure 4.11. Therefore, the most probable kappasquared per se will differ from the most probable kappa-squared at a given efficiency. An algorithm to find the most probable kappa-squared in the second case (at a given efficiency) is briefly as follows (http://www.FRETresearch.org): 1) Choose the a; d pair that best describes the depolarization properties of the actual system. (See Equations 4.12c and 4.12d, and the explanations near these. Note that axial symmetry is assumed. If axial symmetry cannot be justified, see Ref. [11].) j79 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.7 What Is the Most Probable Value for Kappa-Squared in the Dynamic Regime? j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements 2) Because ð1=2Þ a 1 and ð1=2Þ d 1, this pair must lie within one of the regions shown in Figure 4.6 and Table 4.1. Use this figure or table to decide which A A H region 1; 2; . . . ; or 9 applies. Choose B, the number of bins, P P T and vary the bin number i from 1 to B, obtaining bins with relative distance values between rmin þ ði 1Þðrmax rmin Þ=B and rmin þ iðrmax rmin Þ=B, with 1=6 1=6 and rmax ¼ ð3=2Þk2max . rmin ¼ ð3=2Þk2min 3) For n ¼ 1 to B, set QðnÞ to 0. This QðnÞ is the nth component of a B-dimensional array, which will in the end become a histogram approximating the frequency distribution of the relative distance, QðrÞ. 4) Choose N, and in so doing pick N 3 points, varying j, ‘, and m from 1 to N, calculating cos HD ¼ ðj 1Þ=ðN 1Þ, cos HA ¼ ð‘ 1Þ=ðN 1Þ, and W ¼ pðm 1Þ=ðN 1Þ Substitute these into Equation 4.7 to calculate hk2 i values 1=6 and from there relative distance values r ¼ ðð3=2Þhk2 iÞ . Compare each r with the lower and upper boundary of each bin. Place each r in the appropriate bin by adding 1 to each QðnÞ whenever r > rmin þ ðn 1Þðrmax rmin Þ=B and r rmin þ nðrmax rmin Þ=B with 1 n B. 5) Normalize Q by dividing each component by N 3. As a result, the sum of all QðnÞ values will become 1, signifying that the probability that Q has any value equals 1. (For n ¼ 1 to B, QðnÞ ¼ QðnÞ=N 3 .) We have examined graphs of QðrÞ obtained with this algorithm for a large number of points in the plane formed by the depolarization factors a and d, varying these between 0.5 and 1. Results for the most probable kappa-squared are shown in Figure 4.12. The definition of the most probable kappa-squared in Figure 4.8 is that value corresponding to the highest peak in QðrÞ: 2 most probable k2 ¼ r6peak : 3 ð4:18Þ Whenever one or both of the depolarization factors is negative, the most probable kappa-squared is k2P . In the region where both depolarization factors are positive, there is a rather large central region where it is k2L , surrounded by four regions with k2H as the best value and two regions where k2M is the most probable value. The uncertainty in the distance as a result of variations in the orientation factor has two aspects: the most probable kappa-squared may deviate from 2/3, that is, the location of the peak may differ from r ¼ 1, and, the peak may be fairly broad, that is, the 67% confidence interval (CI) may have considerable width (the 67% confidence interval is the range of r-values near the peak where the total QðrÞ adds up to 67%). It is appropriate to call the first aspect a “peak location error” (PLE) and the second a “broad distribution error” (BDE). We note that r ¼ 1 is the relative distance value that corresponds to k2 ¼ 2=3 and, thus, define the PLE as PLE ¼ ð1 rPEAK Þ 100%: ð4:19Þ Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 80 Figure 4.12 Map indicating the most probable kappa-squared in the dynamic regime, where this most probable value is defined as the kappasquared for which the relative distance is the most likely (see text). When one or both of the depolarization factors for the donor or acceptor are negative, k2P is the best. The region where both depolarization factors are positive consists of four regions labeled H, where k2H is the most probable, two regions labeled M where k2M is the best, and one labeled L, where k2L is the most probable kappa-squared value. The border between the H and L regions near the top right corner is well described by a þ d 0:985ad ¼ 1:012. The curved border between H and M on top and L on the bottom, starting at d ¼ 0; a ¼ 0:79 and ending near d ¼ 0:81; a ¼ 1 follows the trend a ¼ 0:79 þ 0:504d4:28 ; and the one with H and M on the right and L on the left is described by d ¼ 0:79 þ 0:504a4:28 . PLE > 0 means that k2 ¼ 2=3 overestimates the most probable distance, and PLE < 0 signifies that this assumption underestimates the distance at the peak. Figure 4.13 shows examples of distributions and PLE values. Our definition of the “broad distribution error” is BDE ¼ 1=2ð67%CIÞ ¼ ðrUL rLL Þ 50%; ð4:20Þ where rLL is the lower and rUL is the upper limit of the 67% CI for QðrÞ. In some cases, the peak is near the center of the confidence interval, but relative distance distributions can also be highly asymmetric with the peak at the upper or lower limit of this interval. Figure 4.14 shows examples. Figure 4.15 shows lines of equal “peak location error” in the ða; dÞ plane. Near a ¼ d ¼ 1 and a ¼ d ¼ 1=2, the PLE is negative, but in the majority of points, the PLE is positive with the most probable distance smaller than the one at k2 ¼ 2=3. A very high positive PLE of about 30% occurs near a ¼ d ¼ 0:96, close to the red line. On the red line, Q has two equally high peaks. The red line is the border between two regions where rpeak is calculated differently. As a result, PLE changes discontinuously at this border. The most dramatic change is at a ¼ d ¼ 0:96 where the PLE is 6%, corresponding to k2H , at the side where the factors are slightly higher than 0.96, and þ30%, corresponding to k2L , at the side where the depolarization j81 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.7 What Is the Most Probable Value for Kappa-Squared in the Dynamic Regime? j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Figure 4.13 Examples of probability density of the relative distance illustrating the definition of the systematic error. The graph on the left is for a ¼ d ¼ 0:73, where the main peak corresponds to k2L with a relative distance smaller than 1 so that PLE is positive (this Q has a secondary maximum corresponding to k2H ). The distribution in the center is for a ¼ 1; d ¼ 0:5 or d ¼ 1; a ¼ 0:5, showing one peak matching k2H ¼ k2M ¼ 2=3, yielding r ¼ 1 and PLE ¼ 0. The graph on the right is for a ¼ d ¼ 1, with a peak at r ¼ 1:07 corresponding to k2H ¼ 1 and a negative PLE. factors are slightly smaller than 0.96. The green lines are also borders between regions where the peak is calculated differently, but with a continuous change in PLE. A large discontinuous change in PLE also implies a fairly broad distance distribution and, therefore, a relatively large BDE. Results for BDE are shown in Figure 4.16. This diagram shows data for the 67% CI, obtained with our program for finding Q (available on the Web site) at any choice for the depolarization factors a and d. To run this program, one must choose an a and a d, a value for B (the number of bins, that is, the number of bars in the histogram approximation for the Q-function), and a value for N (a measure for how many times the relative distance is evaluated; the number of points is N3). After locating the peak (allowing one to confirm the results of Figure 4.12), the CI is obtained by moving away from the peak in both directions while adding the Q-values of the bars in the histogram until 0.67 has been reached. Near the axes, a ¼ 0 or d ¼ 0, the peak is extremely asymmetric with the relative distance at the peak, rPEAK , coinciding with the lower limit of the CI, rLL , at positive a or d, and matching the upper limit rUL at negative a or d, as shown in Figure 4.13. Away from these axes, say for a > 0:1; d > 0:1 or a > 0:1; d < 0:1, or a < 0:1; d > 0:1, or a < 0:1; d < 0:1, the peak is more symmetric and rPEAK is close to the center of the CI. A completely different problem arises near the red borders shown in Figure 4.15. At the red borders, the Q-function has two peaks that are exactly of equal height. For example, at a ¼ d ¼ 0:96, the Q-function has a peak corresponding to hk2 i ¼ k2H ¼ 0:948ðrPEAK ¼ 1:06Þ and an equally high peak for hk2 i ¼ k2L ¼ 0:065ðrPEAK ¼ 0:68Þ. In such cases, the center of the CI should be Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 82 Figure 4.14 Examples of probability density for the relative distance illustrating the definition of the random error. In each, the 67% confidence interval (CI) is shaded dark gray and runs from rLL , the lower limit of the CI, to rUL , the upper limit of the CI. The three graphs have the same scale in r and Q. The one in the center is relatively broad and low. The other two are narrow and high, actually extremely high, as both go to infinity at one point on the interval. The graph on the left is for a ¼ 0:5; d ¼ 0 or d ¼ 0:5; a ¼ 0 with its peak at rUL and a BDE of about 1%. The distribution in the center is for a ¼ d ¼ 1 with its peak near the average of rLL and rUL and a BDE of about 24%. The graph on the right is for a ¼ 1; d ¼ 0 or d ¼ 1; a ¼ 0 with a peak at rLL , and a BDE of about 7%. chosen at the average of the two rPEAK values, and the CI should be built up from there. Examples of graphs for the frequency distributions Q versus the relative distance r are shown in Figure 4.17. 4.8 Optimistic, Conservative, and Practical Approaches For assessing the kappa-squared-induced error in the FRET distance, there is an “optimistic” approach that assuming kappa-squared equals 0.67 introduces little or no error, and there is a “conservative” method based on depolarization factors resulting in a minimum and maximum kappa-squared (with corresponding minimum and maximum distances) without the ability to pinpoint the most probable kappa-squared in this range. The optimistic method is that of Haas et al. [11] and Steinberg et al. [17], and the conservative approach is that of Dale et al. [10]. This classification is an oversimplification, of course, as both the first group and the second group of authors have provided a detailed and versatile discussion of errors j83 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.8 Optimistic, Conservative, and Practical Approaches j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Figure 4.15 Lines of constant “peak location error” are shown with the value of PLE given next to the lines in percent. At the red curves, the relative distance frequency distribution has two equally high peaks. These curves are borders between regions where the most probable kappa-squared is calculated differently, as indicated in Figure 4.12. At one side of a red curve, one of the peaks is highest and on the other side, the other peak is highest. As a result, the PLE changes discontinuously when a red line is crossed. The blue lines are also borders between regions where the most probable kappa-squared is calculated differently, but with a continuous change in the value of the peak and of PLE. resulting from the orientation factor. Nevertheless, neither group has pointed out that there are at least two different aspects associated with the kappa-squaredinduced error: the PLE, introduced in Equation 4.19, and the BDE, introduced in Equation 4.20. For lack of a better name, we would like to call the procedure introduced in the previous section the “practical” approach. Table 4.3 compares the “optimistic,” “conservative,” and “practical” approaches for a range of cases. Note that the PLE in itself is not a problem because when the depolarization factors are known, this error can be accurately predicted using Figure 4.12 and its definition (Equation 4.19). However, the discontinuous jump in the systematic error near ða; dÞ ¼ ð0:96; 0:96Þ may cause serious problems, as a value of 0.96 is experimentally almost undistinguishable from 1 and thus a slight uncertainty in the depolarization factors near this value may cause the PLE to shift from 6 to þ30%. For such high values of a and d, the BDE is also high (see Figure 4.15). Comparing the confidence interval for ða; dÞ ¼ ð1; 1Þ with that for (0.81,0.95) in Figure 4.15 illustrates a problem related to the discontinuity in the PLE: a relatively minor variation in the depolarization factors may cause this interval to shift from one that is centered around r ¼ 1 to one that is centered around 0.85. In Ref. [12], it was assumed that the case a ¼ d ¼ 1 was the worst-case scenario. This is a logical Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 84 Figure 4.16 Regions with lower and higher “broad distribution errors” (BDE), defined in Equation 4.13. A high BDE corresponds to a broad QðrÞ, and low BDE values indicate graphs for QðrÞ with a narrow peak. On the green lines, BDE equals 5%. The long green line connects the points (0.5, 0.31), (0.4, 0.45), (0.3, 0.6), (0.2, 0.66), (0.1, 0.71), (0, 0.78), (0.1, 0.72), (0.2, 0.66), (0.3, 0.58), (0.4, 0.58), (0.54, 0.54), (0.58, 0.4), (0.58, 0.3), (0.66, 0.2), (0.72, 0.1), (0.78, 0), (0.71, 0.1), (0.66, 0.2), (0.6, 0.3), (0.45, 0.4), and (0.31, 0.5). The short green line passes through (0.5, 0.25), (0.33, 0.33), and (0.25, 0.5). In the region between the green lines, BDE is smaller than 5% reaching 0% at a ¼ d ¼ 0. At (0.5, 0.5), BDE ¼ 8%. In between the long green line and the red lines, BDE varies between 5% (on green) and 10% (on red). At (1, 0.5) and (0.5, 1), BDE is about 12% and on the short red curves near these points, RE equals 10%. At (1, 1), BDE ¼ 24%, and BDE decreases with decreasing a and/or d reaching BDE ¼ 10% on the red line connecting (0.45, 1), (0.52, 0.94), (0.66, 0.9), (0.65, 0.83), (0.7, 0.77), (0.74, 0.74), (0.77, 0.7), (0.83, 0.65), (0.9, 0.66), (0.94, 0.52), and (1, 0.45). assumption as a ¼ d ¼ 0 is the best-case scenario and the kappa-squared-induced error gets worse and worse when one moves away from a ¼ d ¼ 0. After Ref. [6] was published, it became possible to generate plots of the frequency distribution of distances and kappa-squared with a few keystrokes on a computer. So, now we must set the record straight: a ¼ d ¼ 1 is not the worst-case scenario as far as the orientation-induced error is concerned in the dynamic regime; it appears that a þ d 0:985ad ¼ 1:012, the red line in Figure 4.15 where the PLE jumps from 30 to 6%, is the worst-case scenario in this regime. 4.9 Comparison with Experimental Results It is imperative to be keenly aware of the assumptions underlying any method one wants to apply. For example, in Refs [1,2], the depolarization factors are not given, j85 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.9 Comparison with Experimental Results j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Figure 4.17 Examples of frequency distributions for the relative distance with the 67% confidence interval (CI) indicated in each as an area between red lines. All graphs have the same vertical scale and the same horizontal scale. The width of each box refers to a relative distance of 1.4, and the height of each box is 9.5 in Q-units. For most choices of a and d, a distribution with one dominant peak is found; but for parameter choices near the red lines in Figure 4.15, more than one equally pronounced peaks may occur, as is shown in the right bottom corner for ða; dÞ ¼ ð0:81; 0:95Þ or (0.95, 0.81). Data for these plots are shown in Table 4.2. The readers will be able to generate their own graphs for any choice of a and d by visiting the Web site http://www.FRETresearch.org. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 86 Table 4.2 Data for Figure 4.17. a or d 1 2 1 2 1 2 1 2 d or a rmaximum rLL rPEAK rUL QðrPEAK Þa) 1 0.953 1.070 0.953 0.953 1.013 1 0.894 1.110 0.958 1.066 1.108 5.613 0.891 1.038 0.986 1.038 1.038 1 0.794 1.134 0.930 0.994 1.056 6.602 0 1.201 0.829 0.949 1.069 1b) 0.891 1.122 0.891 0.891 1.026 1 0.794 1.260 0.887 0.994 1.101 4.295 0.891 1.184 0.929 0.976 1.020 10.077 0 2 1 2 1 2 1 1 3 1 6 1 3 0 4 3 8 3 11 6 4 0.95 0.08 3.379 0 1 0 0.81 rminimum 5 4 5 6 17 12 1 2 1 1 2 1 k2maximum 1 2 1 4 1 3 1 6 0 1 2 1 2 1 k2minimum 0 1.348 0.809 1.051 1.294 2.091 0.702 1.311 0.702 0.848 1.006 2.691 N ¼ 300 and B ¼ 100, but when a or d ¼ 0, data have been calculated analytically (see http://www. FRETresearch.org). a) The frequency distribution of kappa-squared is proportional to that for the relative distance [12] 1=6 according to the relation pðk2 Þ ¼ ð1=4ÞQðrÞ=r5 , with r ¼ ð3k2 =2Þ . b) Mathematically, one can show that the Q-value at the peak is 1, but numerical values depend on B and N. but they should be positive. Therefore, the “practical approach” would suggest that the best kappa-squared value should be k2H , k2L , or k2M. However, in the practical approach, it is assumed that nothing is known about cos HD , cos HA , or W, the variables appearing in Equation 4.10. In the spirit of information theory, it is assumed in this approach that all values of these “hidden variables” are equally probable when no information about them is available. Nevertheless, for the systems in Refs [1,2], information about the relative orientation of donor and acceptor is available: the transition dipoles are essentially perpendicular to the axis of a helix and the angle between the dipoles should depend on the pitch of the helix. This is actually an example of the case where the transfer depolarization is known, within limits, as introduced and analyzed by Dale et al. [10], and in which case equations for kappa-squared have been derived [13]. The geometry of the donor–acceptor pair in Ref. [2] suggests that the best kappa-squared value should be a hybrid between k2H and k2P , k2 ¼ k2HP ¼ ð2=3Þ ð1=3Þa ð1=3Þd þ adcos 2 HT (Equation 38 in Ref. [15]), and that the angle HT between the preferred directions of their transition moments should be equal to 180 r DA =pitch. Substituting this value for HT, and k2 ¼ k2HP , into the expression for the FRET efficiency allows us to model the efficiency versus distance trend. Such an attempt to model the efficiency expected for the system in Ref. [2] is shown in Figure 4.18. j87 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.9 Comparison with Experimental Results Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Figure 4.18 The FRET efficiency versus distance expected for the system of Ref. [2]. Assuming k2 ¼ k2HP ¼ ð2=3Þ ð1=3Þa ð1=3Þd þ adcos 2 QT and QT ¼ 180 r DA =pitch, 0 =pitch (R 0 ¼ with g ¼ r DA =pitch and h ¼ R F€ orster distance when k2 ¼ 2=3) yields a FRET efficiency of the form: E ¼ ½1 ð1= 2Þa ð1=2Þd þ ð3=2Þadcos 2 180 gh6 = g 6 þ ½1 ð1=2Þa ð1=2Þd þ ð3=2Þadcos 2 180 gh6 g. For this graph we chose h ¼ 2, and a ¼ d ¼ 0:8 (solid line) and a ¼ d ¼ 1 (dashed line). The trend described by the solid line is similar to that of the experimental data in Ref. [2]. The kappa-squared in samples 1 and 2 of Ref. [3] is also an example of the case where the transfer depolarization is known. This kappa-squared should be essentially equal to k2A (Equation 4.11a) with values for the depolarization factors close to unity. The history of the kappa-squared fluorescence depolarization relationship is interesting and relevant here. In the late 1970s, when time-resolved fluorescence depolarization was virtually nonexistent, both Dale et al. [10] and Haas et al. [11] realized that information from fluorescence depolarization can be useful for unraveling distance effects from orientation effects in FRET. At that time, both groups had steady-state fluorescence depolarization data in mind. Fairly recently, Dale revisited the kappa-squared fluorescence depolarization relationship [4] and came to the conclusion that this unraveling can be much more effective than previously thought when the full time dependence of fluorescence anisotropy is taken into account. In terms of Figure 4.7, it is fair to say that a glimpse at a short interval on one of the plateaus of anisotropy versus time allows one to put limits on kappa-squared, but a full view of the this curve has the potential to pin down the orientation factor completely. j89 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.9 Comparison with Experimental Results j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements 4.10 Smart Simulations Are Superior Any available information that allows one to exclude certain donor–acceptor orientations will help to narrow down the range of possible kappa-squared values. Simulations can be a powerful tool in this exclusion process. A case in point is the molecular dynamics simulations performed by Lillo et al. [21]. Following the “conservative approach,” these authors found for donors and acceptors at specific sites in a PGK (phosphoglycerate kinase) a fairly large range of possible kappa-squared values and the corresponding donor–acceptor distances, but they noticed that some of the values for HD , HA , and W appearing in Equation 4.10 were inconsistent with the crystal structure of PGK and the excluded volume of the probes at the known sites in PGK. They performed molecular dynamic simulations of kappa-squared utilizing Equation 4.10, measured depolarization factors, the crystal structure of PGK, and the known locations of the donors and acceptors, and found the most probable values of HD , HA , and W, resulting in an improved kappa-squared value and more precise donor–acceptor distances [21]. In the same spirit, Borst et al. built structural models of the FRET-based calcium sensor YC3.60 and noticed that minor structural changes induced by slightly rotating the fluorescent protein around a flexible linker while keeping the same average distance between the donor and the acceptor gave rise to any value of kappasquared between 0 and 3, but a fivefold change in orientation factor (from 0.5 to 2.5) brings only about a 1.3-fold increase in critical distance indicating that the FRET process in YC3.60 is mainly distance dependent [22]. Gustiananda et al. [23] presented FRETresults from an intrinsic tryptophan donor to a dansyl acceptor attached to the Nterminus in model peptides containing the second deca-repeat of the prion protein repeat system from marsupal possum. They used simulations for finding the best kappa-squared in this system and extended their molecular dynamics simulations out to 22 ns to help ensure adequate sampling of the dansyl and tryptophan ring rotations. They found good agreement of the simulated kappa-squared value with 2/3, except at the lowest temperatures [23]. Deplazes et al. performed molecular dynamics simulations of FRET from AlexaFluor 488 donors to AlexaFluor 568 acceptors [24]. In their system, the isotropic dynamic condition was met, meaning that all possible orientations of the transition moments of donor and acceptor and of the line connecting their centers are equally probable and sampled within a time short compared to the inverse transfer rate. The frequency distribution (Figure 4.19) of kappa-squared from the simulation data showed excellent agreement with the theoretical distribution [12]: 8 pffiffiffi 1 > > 0 k2 1; < pffiffiffiffiffiffiffi2 ln ð2 þ 3Þ; 2 3k pffiffiffi pðk2 Þ ¼ ð4:21Þ 1 2þ 3 > > : pffiffiffiffiffiffiffi ln pffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffi ; 1 k2 4: 2 3k2 k2 þ k2 1 Their results show that even in their simple situation, simulations lasting longer than 200 ns would be required to accurately sample the fluorophore separations and kappa-squared if only a single donor–acceptor pair had been included. Many aspects Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 90 Figure 4.19 The probability density pðk2 Þ versus k2 , as described by Equation 4.21, in the dynamic regime for the case a ¼ d ¼ 1. of FRET were simulated in this study, including frequency distributions of relevant angles, donor–acceptor distance, and FRET efficiency. As expected, very low correlation was found between donor–acceptor distance and orientation factor [24]. VanBeek et al. did find such a correlation in a molecular dynamics simulation of a coumarin donor and an eosin acceptor, both attached to HEWL (hen egg-white lysozyme) [25]. In the dynamic regime, it is implicitly assumed that kappa-squared is independent of the donor–acceptor distance. (In the static regime, an indirect correlation between distance and kappa-squared is expected, as discussed near Equation 4.31). The correlation between orientation and distance in the molecular dynamics study of Vanbeek et al. is quite strong and involves both the sign and the magnitude of kappa (k ¼ cos qT 3cos qA cos qD , the square of which is given in Equation 4.1, where the angles are also defined). This correlation is illustrated in Figure 4.20, which is a modification of Figure 6 of Ref. [25], graciously made available for this chapter by Dr. Krueger. An additional advantage of molecular dynamics simulations is that no assumptions about timescales need to be made, whereas in the interpretations of FRETexperiments, the results do depend on whether the system in is the dynamic or static regime. Note that the FRET efficiency also shows a relationship with kappa and the donor– acceptor distance in this illustration. The kappa-squared concept is based on the ideal dipole approximation that is known to fail when molecules get “too close” to each other. Mu~ noz-Losa et al. performed molecular dynamics simulations to find out how “too close” should be defined [26]. They showed that the ideal dipole approximation performs well down to about a 2 nm separation between donor and acceptor for the most common fluorescent probes, provided the molecules sample an isotropic set of relative orientations. If the probe motions are restricted, however, j91 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.10 Smart Simulations Are Superior j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Figure 4.20 Modification of Figure 6 of Ref. [25]: a scatter plot of the donor–acceptor distance against kappa showing the correlation between this distance, kappa, and the FRET efficiency. The color code on the right is for the efficiency. This graph has been prepared by Dr. Brent Krueger. this approximation performs poorly even beyond 5 nm. In the case of such restricted motion, FRET practitioners should worry not only about kappa-squared but also about the failure of the ideal dipole approximation [26]. In a more recent paper from the same laboratory, an improved construction of experimental observables from molecular dynamics sampling has been proposed [27]. Hoefling et al. have introduced a similar analysis [28]. 4.11 Static Kappa-Squared We begin our consideration of the impact of molecular motion during the excited state on FRET by considering how the rate of FRET is influenced by the separation between donor and acceptor (rDA) as well as by the orientation of donor and acceptor dipoles relative to each other (k2). The rate of energy transfer by FRET, kT, is dependent on the inverse sixth power of the separation between donor and acceptor [29,30]: 1 R0 6 kT ¼ ; ð4:22Þ t0D r DA Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 92 where t0D is the fluorescence lifetime of the donor molecule and R0 is the F€orster distance, the separation at which 50% of the donor excitation events result in energy transfer to the acceptor. Furthermore, the R0 value used for any specific donor– acceptor FRETpair always assumes that the dipole–dipole coupling orientation factor (k2) will have a value of 2/3, but in reality it can have any value from 0 to 4 in biological experiments, and can be expressed as [12,17] k2 ¼ 1 þ 3x 2 z2 : ð4:23Þ This is Equation 4.3 with the abbreviations z ¼ jcos vj and z ¼ jcos qD j, where qD is the angle between the donor emission dipole orientation and the donor–acceptor separation vector and v is the angle between the donor electric field vector at the acceptor location and the acceptor absorption dipole orientation. For a typical donor fluorophore with a fluorescence lifetime of 3 ns, more than 99% of excited fluorophores have returned to their ground state within approximately 15 ns, that is, within about five lifetimes. It is therefore reasonable to consider (i) if the separation rDA can change during this period, (ii) if the position of the acceptor relative to the donor emission dipole orientation can change during the excited state and therefore (except for change only along the separation vector) the value of qD , and, pari passu, that of the field vector at the acceptor location, (iii) if the orientation of the donor emission dipole changes, and thus again the value of qD changes, and finally (iv) if the angle between the donor electric field vector at the acceptor location and the acceptor absorption dipole orientation (v) changes within this 15 ns period. Changes in rDA and/or in qD can be caused by significant lateral motion of the acceptor fluorophore relative to the position of the donor fluorophore. Thus, our first consideration should be how far a fluorophore can move by diffusion in 15 ns? Diffusion is a function of the mass of the molecule, its hydrodynamic shape, and the temperature, as well as the viscosity of the milieu. Assuming a temperature of 20 C and an essentially aqueous local environment, a small fluorophore may have a diffusion coefficient between 100–1000 mm2/s, while a larger fluorophore like GFP will have a diffusion coefficient of 70 mm2/s. Under these conditions, one might expect that a free fluorophore could diffuse a distance between 1.4–5.5 nm during a 15 ns excited state. Clearly, such motion could influence the effective value of both rDA and qD in a FRET experiment. In practice, however, most donor fluorophores will return to the ground state in a much shorter time span, with a median value (50% of excited states lost) of ðln 2Þ t0D , in this instance 2 ns, effectively limiting the distance that most molecules (about 80%) can diffuse by up to 0.5–2.0 nm away from their original location. Furthermore, when one considers that fluorophores used in biological FRET experiments are typically coupled to much larger molecules such as protein complexes or nucleic acids with much smaller diffusion coefficients in aqueous solution, and even smaller in cell cytoplasm in which the local viscosity for these large molecules is much higher than that of water, it is typically assumed that lateral motion during the excited state will be so limited that it will not be responsible for any alterations in the rDA or qD values for a specific pair of molecules tagged with donor and acceptor fluorophores. j93 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.11 Static Kappa-Squared j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements In addition to lateral diffusion, another type that must be considered is that due to molecular rotation. Specifically, we will consider if donor and acceptor fluorophores can rotate during the excited state, and if so, the impact of this rotational motion on the values of qD and v, and thus on FRET. Molecular rotation is typically parameterized by a rotational correlation time ðtrot Þ, the average time that it takes a molecule to rotate 1 rad around a specific axis. For spherical molecules, the rotational correlation time will be the same around all three axes. Nonspherical molecules can have different diffusion coefficients for each principal axis of rotation. These are identical for spherical molecules, leading to a monoexponential decay of the emission anisotropy of a fluorescent probe rigidly associated with this structure, with a rotational correlation time proportional to the inverse of the diffusion coefficient. For ellipsoids of revolution, there are two different ones, one for the axis of symmetry and another for the two principal axes perpendicular to this, leading, in general, to a triple exponential decay of the anisotropy with three different correlation times: one associated with rotation of the unique axis being proportional to the inverse of the diffusion coefficient of the axis of symmetry, and two associated with both this rotation and that of the equivalent perpendicular principal axes and differing in the contributing weights of their summed diffusion coefficients defining the inverse of the correlation times. Rotational correlation times of fluorescent molecules can be measured experimentally by monitoring the decay of fluorescence anisotropy as a function of time after a transient excitation pulse [31]. In the absence of homo-FRET, this decay is primarily caused by molecular rotation. By fitting the anisotropy decay to a triexponential model, the rotational correlation time or times can be estimated. In the case of ellipsoids of revolution, the general form of the time-dependent decay of the fluorescence anisotropy, r(t), is given by (compare Equation 4.5) r ð tÞ ¼ r 0 i¼3 X ai et=troti ; ð4:24Þ i¼1 where r0 is the limiting anisotropy, the initial anisotropy at the instant of photoexcitation prior to any rotational depolarization, ai is the amplitude of the ith decay component, and troti is the rotational correlation time of the ith decay component. The contribution of all three components to the decay depends on the orientations of the absorption and emission transition dipoles in the molecular frame: each one alone or any combination of pairs may appear and, in addition, the anisotropy may decay monotonically from either positive or negative values or start at zero or a positive or negative value, then increase or decrease before changing direction toward zero at long enough times, or even cross zero before turning over and moving toward zero [32]. In practice, differences in rotational correlation times for the three axes for most fluorophores are hard to experimentally distinguish, and, more typically, the monoexponential anisotropy decay will statistically adequately fit the anisotropy decay data, that is, when rotational diffusion coefficients are similar enough, the extent of similarity required depending on the level down to which the Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 94 anisotropy is accurately recovered [33]: r ðtÞ ¼ r 0 et=trot : ð4:25Þ Here the value of trot is a function of the solution viscosity (g), temperature (T), and the molar volume of the rotating molecule (V) [33]: trot ¼ gV ; RT ð4:26Þ where R is the gas constant. For example, small fluorophores, such as fluorescein (332.31 g/mol), will have a rotational correlation time of 140 ps in water at room temperature, while a large 28 000 Da fluorophore such as Venus (a yellow GFP derivative) has a rotational correlation time of 15 ns under the same conditions [31], presumably because the volume of Venus is approximately 100 times greater than the fluorescein. When a population or randomly oriented fluorophores (isotropic) are photoselected using a linearly polarized light source, the highest anisotropy value theoretically possible (fundamental anisotropy) is 0.4 with onephoton excitation and 0.57 with two-photon excitation [31]. In practice, other factors can reduce the value of the initial anisotropy value at time ¼ 0. Thus, the limiting anisotropy measured in a time-resolved anisotropy measurement is usually smaller than the fundamental anisotropy expected from theory. With time, measured anisotropy values for fluorophores in solution that are free to rotate in any direction will decrease as a single exponential with an asymptote at 0. This value indicates the point where all remaining molecules in the excited state are randomly oriented. The speed of this orientational randomization is parameterized by the rotational correlation time. For a system decaying as a single exponential, this occurs at 5X, the rotational correlation time. Thus, for a small molecule like fluorescein, nearcomplete orientational randomization can occur within 700 ps, well within the excited-state lifetime of fluorescein (4.1 ns). In contrast, for Venus under the same conditions, this would require 75 ns, much longer than its lifetime of 3 ns. As mentioned above, most of the excited donor fluorophores in a FRET experiment will return to the ground state in a much shorter time span, with a median value of ðln 2Þ t0D , (for Venus, 2 ns). With a rotational correlation time of 15 ns, free Venus is only expected to rotate 11 in 2 ns. Furthermore, Venus will rotate even slower when attached to another protein, or if situated in the more viscous cytoplasm found in cells. Thus, Venus is not expected to rotate much during its excited state. In contrast, a small fluorophore like fluorescein may be able to rotate during its excited state. Thus, when considering the value of k2 to use in interpreting a FRET experiment, it is important to note that the values qD and v may be average values over many possible angles when small fluorophores are used as FRET donors and acceptors, while the values for qD and v may be static for any particular donor– acceptor pair composed of fluorescent protein donors and acceptors. At this point, it is worth noting and yet again emphasizing that the 2/3 value for k2, so ubiquitously used in FRET experiments, is based on two assumptions: (i) That, bar a fortuitously occurring set of static relative orientations leading to this value, qD and v have random values (i.e., they come from isotropic distributions). (ii) That the values of j95 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.11 Static Kappa-Squared j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements qD and v are changing rapidly relative to the fluorescence lifetime (dynamic). From the above calculations, it should be clear that these assumptions (isotropic dynamic) might be valid for some FRETexperiments using small fluorophores like fluorescein that can rotate rapidly, provided that they are attached via flexible-enough chains linking them to the protein backbone and/or donor and acceptor to different smallenough segmentally flexible units in the protein, but are not generally expected to be valid for FRET experiments using fluorescent proteins as donors and acceptors because being bound into the alpha-helical backbone of the protein, they hardly rotate at all during their fluorescent lifetimes (static, with the possible exception that they may be bound to separate subunits of the protein, one or both of which may exhibit rapid segmental flexibility). What k2 value should be used in a FRET experiment if one assumes that the values of qD and v are randomly selected from isotropic populations, but the donor and acceptors are in the static regime, that is, they are hardly rotating during the excitedstate lifetime of the donor? Steinberg et al. [17] have shown that in the static regime, hk2 i for an isotropic population varies with separation in a sigmoid fashion, starting essentially at zero at very low distance, eventually leveling off at a value of 2/3 at very large distances [12]. Dale has derived an approximation for an effective-kappasquared value for use in the static regime: hk2 ief f 2/3ð1 hE iÞ [34]. Recently, Monte Carlo simulations were used to address this same issue [35]. This study confirmed Steinberg’s finding that no single value of hk2 i can be used to predict the energy transfer behavior of a static population (and is in good agreement with the Dale effective kappa-squared approximation [34]), rather it was found that a hk2 i value must be calculated from the random values of qD and v on a FRET pair by FRET pair basis for each pair in the population. What emerged from this study is that even for a population that has a homogeneous separation that strongly favors energy transfer by FRET (rDA < R0), because the most probable value of hk2 i for an isotropic population is zero [12], a large fraction of FRET pairs in a population will only transfer a negligible fraction of their excitation energy by FRET, and the population behavior will be heterogeneous with some FRET pairs having very efficient transfer and some having none (k2 ¼ 0) or essentially none (k2 near 0). With respect to FLIM measurements of donor lifetimes from an isotropic static population of donors and acceptors, a simple single-exponential decay is expected only if the rDA value is much larger than the R0 value (approximately no FRET). In this case, the simple lifetime decay would be the same as the decay of donor alone. If the rDA value is short enough to support a significant amount of FRET, a multiexponential decay is expected even when only a single fixed rDA value is present in the population. In this instance, the average FRET efficiency calculated from the multiexponential decay may be smaller than that obtained through steady-state intensity measurements for two reasons. First, low kappa-squared values are more common than high values, and second, because a fraction of donor–acceptor pairs with relatively large kappasquared values will exhibit such efficient transfer (approaching unity) that the donor lifetimes will be beyond the resolution of the measurement and so not appear in the decay curve. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 96 Is there experimental evidence for static FRET behavior in experiments with fluorescent protein donors and acceptors? Specifically, for FRET in the static isotropic regime, we expect to observe (i) a complex multiexponential donor lifetime, even for a homogeneous population of FRET pairs and (ii) a large fraction of FRET pairs in the population should fail to transfer energy by FRET because of the prevalence of low k2 values expected in an isotropic population and the absence of appreciable rotational motion during the excited state, even when separation between donors and acceptors are short. In Figure 4.21a, a three different constructs are depicted, each engineered to express in cells a Cerulean [36] FRET donor (a blue GFP derivative) covalently attached to a Venus [37] acceptor (a yellow GFP derivative) via a 5-, 17-, or 32-amino acid linker. These constructs are called C5V, C17V, and C32V, respectively [38,39]. As a negative FRET control, a single point mutation was introduced into Venus at Y67 (from Y to C) to form “Amber,” a protein that is thought to have the same structure as Venus, but cannot form the Venus fluorophore and does not act as a dark absorber in the region of Cerulean fluorescence [40]. This Amber mutation was then used to create three more constructs; C5A, C17A, and C32A. While the Cerulean decays of C5A, C17A, and C32A are indistinguishable (Figure 4.21b), the decays of C5V, C17V, and C32V were all faster than the Cerulean–Amber constructs, with C5V having the fastest decay, and C32V having the slowest. Using these Cerulean decays in the presence and absence of acceptor (Venus), C5V with its short 5-amino acid linker had the highest average FRET efficiency [(43 2)%], the FRET efficiency of C17V was intermediate [(38 3)%], and C32V, with the longest linker separating the donor from the acceptor, had the lowest FRET efficiency [(31 2)%] [38]. Note that C5V, C17V, and C32V all have complex decays that are clearly not single exponential, even though every expressed molecule in the population should have one Cerulean donor covalently attached to one Venus acceptor. These complex multiexponential fluorescence lifetime decays for donor covalently attached to acceptors suggest that the underlying distribution of FRET efficiencies in these populations is heterogeneous. While this complex decay behavior is consistent with the first prediction of FRET in the static isotropic regime, somewhat awkward is the observation that the lifetimes of the three corresponding Cerulean–Amber constructs did not decay as a purely single exponential as the F€ orster theory predicts for donor-only constructs. This might arise from more complicated photophysics for fluorescent protein fluorophores, perhaps indicating multiple excited states for these fluorophores. While such complicated donor-alone decay behavior is problematic, it is quite typical for decays of isolated fluorescent proteins, not to say ubiquitous, and has been observed in experiments measuring FRET between spectral variants of many different fluorescent proteins [41]. Regardless, to test the second prediction of FRET in the static isotropic regime, an analysis method is needed that can account for the complex decay behavior of the donor-alone. To look for a fraction of molecules in a population that does not undergo energy transfer, the data plotted in Figure 4.21b were transformed and replotted as the time-resolved FRET efficiency (TRE) (Figure 4.21c) This transformation involves j97 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.11 Static Kappa-Squared j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements Figure 4.21 (a) Cartoons depicting the FRETpositive protein constructs C5V, C17V, and C32V and their FRET-negative analogues C5A, C17A, and C32A, where C stands for Cerulean (donor), V for Venus (acceptor), and A for Amber (VenusY67C), a nonabsorbing Venus with a single-point mutation that prevents chromophore formation. The number between C and V and C and A denotes the number of amino acids in the linker connecting them. (b) Donor fluorescence intensity, IDA, versus time after donor excitation in the presence of energy transfer to Venus for C5V, C17V, and C32V, and intensity, ID, versus time in the absence of energy transfer for C5A, C17A, and C32A. (c) Experimental TRE versus time for C5V, C17V, and C32V compared to C5A, C17A, and C32A. (d) Theoretical TRE versus time based on Equation 4.28 in which IDA(t) is calculated assuming that the excited donor is characterized by a monoexponential decay with time constant equal to the appropriate average of the measured lifetimes for a population of donor–acceptor pairs over which kappasquared is randomly distributed in the static limit, with choices for the relative distance (R0/ rDA with R0 defined for kappa-squared ¼ 2/3) that yield a strong resemblance to the experimental curves in panel (c). Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 98 calculating the time-dependent change in FRET efficiency normalized to the fluorescence lifetime decay of the donor: TREðtÞ ¼ I D ðtÞ I DA ðtÞ ; ID ðtÞ ð4:27Þ where ID(t) is the fluorescence decay of the donor-alone and IDA(t) is the fluorescence decay of the donor in the presence of acceptor. Note that ID(t) does not have to be monoexponential, it could just as well have a more complex decay resulting from the sum of multiple excited states. Similarly, IDA(t) can also be a complex decay resulting from multiple decay components, but including a component, or components, representing energy transfer by FRET from the donor (or multiple donor excited states) to an acceptor (or multiple acceptor ground states). If every donor or donor excited state that registers in the IDA(t) determination undergoes FRET, the TRE curves will start at a value of 0 at time 0 and eventually asymptote at a TRE value of 1. In contrast, if some donors or donor excited states never transfer energy by FRET, as predicted for energy transfer in the static isotropic regime, the TRE curve will still start at a value of 0 at time 0, but appear to asymptote to a TRE value that is less than 1. This difference represents the fraction of molecules in the population that do not transfer energy by FRET and/or that transfers with very high efficiency and is not detected. In Figure 4.21c, we can see that the TRE curves for the decay data presented in panel (b) for C5V (and C5A), C17V (and C17A), and C32V (also C32A) all seem to asymptote to a value that is between 0.71 and 0.73, indicating that for these constructs approximately 27–29% of the donors either do not transfer energy and/or very efficiently transfer it by FRET (or any other additional mechanism). This type of behavior is consistent with the predictions of FRET in the static isotropic regime, but other sources of population heterogeneity [35] may also participate in producing a TRE curve asymptote of less than 1. The main advantage of TRE analysis over directly examining fluorescence lifetime decay curves is that TRE analysis facilitates discriminating between population FRET behavior in the dynamic and static regimes. If all donor–acceptor pairs in the sample behave similarly and are expected to have the same overall efficiency, the TRE curve will be 1 minus a single exponential. In contrast, if a distribution of efficiency values is present in the system, a sharp deviation of this trend will be seen. It is expected that a single-exponential TRE curve could be a signature for the dynamic regime, whereas the static regime may be characterized by a more complex TRE curve appearing to asymptote to a value less than 1. In the static isotropic regime, theory predicts that the TRE curve should follow the following trend: pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffi ð1 ð1 ð1 erf ð1 þ 3x 2 Þy p 2 2 ð4:28Þ TRE ¼ 1 dx dzez ð1þ3x Þy ¼ 1 dx pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; 2 ð1 þ 3x 2 Þy 0 0 0 where x and z are introduced in Equation 4.16, erf denotes the error function, and y is given by 3 R0 6 t y¼ ; ð4:29Þ 2 r DA t0D j99 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.11 Static Kappa-Squared j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements where t is the time, t0D is the average donor lifetime in the absence of transfer, R0 is the F€orster distance when k2 ¼ 2=3. The rDA values estimated by TRE analysis assuming a static isotropic regime for C5V, C17V, and C32V (5.0, 5.3, and 5.5 nm, respectively) are lower than the rDA values estimated from the average efficiency and fluorescence lifetime decay analysis assuming a dynamic isotropic regime (5.7, 5.9, and 6.2 nm, respectively). This is expected because a large fraction of the FRETpairs in an isotropic static regime population will have k2 values close to zero. It is clear that the experimental TRE data (Figure 4.21c) are not in perfect agreement with the theoretical TRE results based on Equation 4.28 (compare Figure 4.21c and d). While the basis of these small discrepancies is not known, we speculate that fluctuations in the separation between donors and acceptors, or deviations from a purely isotropic distribution of qD and v angles, which are not taken into account in Equation 4.28, may explain this discrepancy. Regardless, it is quite remarkable that with only one adjustable parameter, 3R60 = 2t0D r 6DA , the agreement between theory and experiment is as good as it is, clearly indicating, we believe, that the static regime character of kappa-squared is the major reason for why the time-resolved efficiency for C5V, C17V, and C32V deviates so dramatically from a single exponential rising from 0 to 1. If it is known that a population of FRET pairs are in the static isotropic regime, with a few assumptions it is also possible to estimate the donor–acceptor separation from experimentally measured hk2 i values using as our starting point an estimate of the average kappa-squared in the static regime introduced by Steinberg et al. [17]: 0 ð3=2Þhk2 iR ; 6 2 ð3=2Þhk iR þ r 6 6 hE i ¼ 0 ð4:30Þ DA 0 ¼ R0 is the F€ The angle brackets in this equation denote an average, R orster distance when k2 ¼ 2=3, and r DA is the donor-acceptor distance. Steinberg et al. have shown, in a graph, that hk2 i varies with distance in a sigmoid fashion in the static regime, starting essentially at zero at very low distance, then rising slowly until about r DA ¼ 2=5R0 , where hk2 i starts to increase more strongly with increasing distance until about r DA ¼ 7=5R0 , where hk2 i begins to level off reaching 2/3 at very large distances [12]. Between r DA ¼ 2=5R0 and r DA ¼ 7=5R0 , hk2 i varies linearly with distance and is approximately equal to 2=3ðr DA =R0 ð2=5ÞÞ [12]. For example, the distances between the Cerulean and Venus fluorophores in C5V, C17V, and C32V most likely fall in this range between 0:4R0 and 1.4R0 (2.2–7.7 nm). Substituting hk2 i ¼ 2=3ðu 2=3Þ (with u r DA =R0 ) into (4.23) yields Equation 4.31 for u: 1 hE i 2 u : u6 ¼ ð4:31Þ 5 hE i Solving this equation numerically using the measured average efficiencies and R0 ¼ 5:4 nm, the estimated rDA values are found to be 5.1, 5.4, and 5.8 nm for C5V, C17V, and C32V, respectively, in excellent agreement with distance estimates derived from TRE analysis (5.0, 5.3, and 5.5 nm respectively), and with r DA values obtained after substituting Dale’s approximation, hk2 i ð2=3Þð1 hE iÞ [34], into Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 100 equation 4.30 (5.2, 5.4, and 5.8 nm respectively). Interestingly, this substitution qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 0 6 ð1 hE iÞ2 =hE i. leads to r DA R With regard to FRET in the static regime, it is important to realize that it is possible to be in the static regime even when the FRET donor and acceptor used in an experiment are small fluorophores like fluorescein. Clearly, experimental factors such as high viscosity or short rigid linkers can restrain the motion of a small fluorophore. Similarly, if fluorescent protein donors and acceptors are attached to interacting proteins via a short rigid linker, the values of qD and v, and thus k2, may be fixed and identical for every FRET pair in the population. If this is the case, FLIM– FRET analysis will reveal a simple exponential decay that is faster than the decay of the donor alone, and TRE curves will asymptote from 0 to a value of 1. In this case, we are still in the static regime, but since the kappa-squared value is unique, the isotropic assumption is obviously no longer valid. 4.12 Beyond Regimes It is possible that the average rate of transfer is on the same order of magnitude as a dominant rate of rotation for the donor or acceptor. In this case, the system is neither in the dynamic regime nor in the static regime. Molecular dynamics simulations are extremely useful in this intermediate regime [19–23]. Analysis is still possible by building mathematical models based on the idea that a system of donors and acceptors undergoing translational and/or rotational motion during the transfer time (inverse of the average transfer rate) can be described as a collection of states with transitions between them [42]. These states can be visualized as snapshots: at a certain moment, a donor is excited and has a particular orientation, while the acceptor has another orientation. This donor–acceptor pair is then in a D A state. A little later the donor or acceptor changes its orientation, that is, a rotational transition to another D A state has occurred. FRET corresponds to a transition to a DA state. A systematic description of such time developments implies selecting a representative set of orientation states, evaluating kappa-squared values, and identifying transfer rates and rates of rotation. This approach leads to a matrix equation for which the eigenvectors and eigenvalues must be found, so that intensities and anisotropies can be calculated [42]. The following example illustrates this method. A donor and acceptor are at a fixed distance r DA from each other. The acceptor’s absorption moment has an isotropic degeneracy. The donor’s emission moment is linear and can only have two orientation states: parallel to the “connection” line (line connecting the centers of donor and acceptor) or perpendicular to it. The rate of rotation of 6 this moment is 1=tR . The FRET rate is ð3=2Þk2 t1 0D r , where t0D is the fluorescence lifetime of the donor in the absence of FRET and r is the relative distance (r DA divided by the F€orster distance if kappa-squared would be equal to 2/3). In this example, k2 equals either 4=3 or 1=3: 4=3 when the donor is in the “parallel” state with its moment parallel to the connection line and 1=3 when the donor is in the j101 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 4.12 Beyond Regimes j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements “perpendicular” state with its moment perpendicular to this line. I DA , the fluorescence intensity of the donor in the presence of acceptor after excitation with a very short pulse of light, is proportional to y== þ y? . Here, y== is the fraction of the donors with its moment parallel to the connection line and y? is the fraction with this moment perpendicular to the connection line. For ID, the fluorescence intensity of the donor in the absence of FRET, y== ¼ y? ¼ 1=2 at all times, but for IDA, y== ¼ y? ¼ 1=2 only at time zero when the system is excited by the flash, whereas at later times y== > y? until they both decay to zero at times much larger than t0D . The rate equation for this example is d dt y== y? ¼ 1 t1 t1 0D þ tR ½1 þ ð8=3Þx R 1 1 t1 t þ t R 0D R ½1 þ ð2=3Þx y== ; y? ð4:32Þ where the differentiation is with respect to the time t and x ¼ ð3=4Þr6 t1 0D tR . The time-resolved efficiency TRE (defined in Equation 4.27) can be calculated from the solution of (4.32) in terms of the two eigenvectors with the initial condition y== ¼ y? ¼ 1=2 and for this example reads (see FRETresearch.org for details) ! pffiffiffiffiffiffiffi2ffi 1 1 TRE ¼ 1 1 pffiffiffiffiffiffiffiffiffiffiffiffiffi eðt=tR Þ 1þð5=3Þxþ 1þx 2 1 þ x2 ! pffiffiffiffiffiffiffi2ffi 1 1 1 þ pffiffiffiffiffiffiffiffiffiffiffiffiffi eðt=tR Þ 1þð5=3Þx 1þx : 2 1 þ x2 ð4:33Þ The special cases for this example are as follows: No FRET with x ¼ 0 and TRE ¼ 0. The static regime with x ! 1, tR ! 1, while x=tR remains at ð3=4Þr6 t1 0D , 2r6 t1 ð1=2Þr6 t1 0D t 0D t yielding TRE ¼ TRESTATIC ¼ 1 ð1=2Þe ð1=2Þe . The dynamic regime with x ! 0, tR ! 0, while x=tR remains at ð3=4Þr6 t1 0D , ð5=4Þr6 t1 0D t yielding TRE ¼ TREDYNAMIC ¼ 1 e . 4.13 Conclusions In FRET situations where the transition moments of donor and acceptor are isotropically degenerate or reorient rapidly and completely within a time comparable to the lifetime of the excited donor state in the presence of acceptor, one can be certain that kappa-squared equals 2/3. Often this simplification is not warranted. However, we have indicated which methods can be utilized to diagnose the potential problems caused by the orientation factor, which alternative value can be used if the experimental conditions allow one to find an average or actual kappa-squared value, and what can be done in cases where an average value is poorly defined. The concept Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 102 of time-resolved FRET efficiency (Equation 4.27) can be useful, especially when combined with the idea of relaxation of the probability density pðk2 Þ: In the dynamic regime, pðk2 Þ relaxes from a relatively broad distribution to a narrow peak within a time much shorter than the lifetime of the excited state of the donor, whereas in the static regime this relaxation takes much longer than this lifetime. Acknowledgments We wish to thank Dr. Bob Dale for many helpful suggestions and stimulating discussions. Bob also made us aware of important papers we had overlooked. Drs. Manual Prieto and David Lilley mentioned relevant papers as well. We thank them for that. Dr. Klaus Suhling gave us useful ideas to improve the explanation of relevant issues. Dr. David Piston suggested to add electric field lines to Figure 4.5. We wish to acknowledge Dr. Paul Blank for stimulating discussions on strategies for fitting TRE decays to characterize separation distance in the static isotropic regime and Dr. Brent Krueger for designing Figure 4.20 especially for us, allowing us to use it in this chapter, and for stimulating discussions about kappa-squared and molecular dynamics simulations. We are indebted to Dr. Phil Womble for writing a program, allowing us, with help from Sandeep Kothapalli, to obtain some preliminary data for the preparation of Figure 4.12. We thank Sarah Witten Rogers for valuable help in calculating frequency distributions for the relative distance. S.S.V. was supported by the intramural program of the National Institute of Health, National Institute on Alcohol Abuse and Alcoholism, Bethesda, MD. References 1 Lewis, F.D., Zhang, L., and Zuo, X. (2005) Journal of the American Chemical Society, 127, 10002–10003. 2 (a) Iqbal, A., Arslan, S., Okumus, B., Wilson, T.J., Giraud, G., Norman, D.G., Ha, T. and Lilley, D.M.J. (2008) Proceedings of the National Academy of Sciences, 105, 11176–11181; in a more recent paper from the same lab, the effect of a flexible tether is discussed: (b) Urnavicius, L., McPhee, S.A., Lilley, D.M.J., and Norman, D.C. (2012) Biophysical Journal, 102, 561–568. 3 Ranjit, S., Gurunathan, K., and Levitus, M. (2009) The Journal of Physical Chemistry B, 113, 7861–7866. 4 Dale, R.E. (2002) Presentation at the American Biophysical Society Meeting, and private communication. 5 Dale, R.E. (1979) Conference Digest, vol. 2, 6 7 8 9 10 3rd Conference on Luminescence, Szeged, Hungary. Dale, R.E. (2013) private communication. Tweet, A.G., Bellamy, W.D., and Gaines, G.L., Jr. (1964) The Journal of Physical Chemistry, 41, 2068–2077. Loura, L.M.S., Federov, A., and Prieto, M. (2000) Biochimica et Biophysica Acta, 1467, 101–112. Loura, L.M.S. and Prieto, M. (2000) The Journal of Physical Chemistry B, 104, 6911–6919. Dale, R.E., Eisinger, J., and Blumberg, W.E. (1979) Biophysical Journal, 26, 161–194 (with Appendix B corrected in Biophysical Journal (1980) 30, 365). j103 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 4 Optimizing the Orientation Factor Kappa-Squared for More Accurate FRET Measurements 11 Haas, E., Katchalski-Katzir, E., and 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 Steinberg, I.Z. (1978) Biochemistry, 17, 5065–5070. van der Meer, B.W. (2002) Reviews in Molecular Biotechnology, 82, 181–196. B€ urkli, A. and Cherry, R.J. (1981) Biochemistry, 20, 138–145. Corry, B., Jayatillaka, D., Martinac, B., and Rigby, P. (2006) Biophysical Journal, 93, 1032–1045. van der Meer, B.W. (1999) Resonance Energy Transfer (eds D.L. Andrews and A.A. Demidov), John Wiley & Sons, Ltd, Chichester, pp. 151–172. Loura, L.M.S., Carvalho, A.J.P., and Ramalho, J.P.P. (2010) Journal of Molecular Structure: THEOCHEM, 946, 107–112. Steinberg, I.Z., Haas, E., and KatchalskiKatzir, E. (1983) Time-Resolved Fluorescence Spectroscopy in Biochemistry and Biology (eds R.B. Cundall and R.E. Dale), Plenum Press, New York, pp. 411–450. Dos Remedios, C.G., and Moens, P.D.J. (1999) Resonance Energy Transfer (eds D.L. Andrews and A.A. Demidov), John Wiley & Sons, Ltd, Chichester, pp. 1–64. Wagner, R., Podesta, F.E., Gonzalez, D.H., and Andreo, C.S. (1988) European Journal of Biochemistry, 173, 561–568. Lankiewicz, L., Malicka, J., and Wiczk, W. (1997) Acta Biochimica Polonica, 44, 477–490. Lillo, M.P., Beechem, J.M., Szpikowska, B.K., Sherman, M.A., and Mas, M.T. (1997) Biochemistry, 36, 11261–11272. Borst, J.W., Laptenok, S.P., Westphal, A.H., K€ uhnemuth, R., Hornen, H., Visser, N.V., Kalinin, S., Aker, J., and Visser, A.J.W.G. (2009) Biophysical Journal, 95, 5399–5411. Gustiananda, M., Liggins, J.R., Cummins, P.L., and Gready, J.E. (2004) Biophysical Journal, 86, 2467–2483. Deplazes, E., Jayatilaka, D., and Corry, B. (2011) Physical Chemistry Chemical Physics, 13, 11045–11054. VanBeek, D.B., Zwier, M.C., Shorb, J.M., and Krueger, B.P. (2007) Biophysical Journal, 92, 4168–4178. Mu~ noz-Losa, A., Curutchet, C., Krueger, B.P., Hartsell, L.R., and Mennuci, B. (2009) Biophysical Journal, 96, 4779–4788. 27 Speelman, A.L., Mu~ noz-Losa, A., Hinkle, 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 K.L., VanBeek, D.B., Mennuci, B., and Krueger, B.P. (2011) Journal of Physical Chemistry A, 115, 3997–4008. Hoefling, M., Lima, N., Haenni, D., Seidel, C.A.M., Schuler, B., and Grubm€ uller, H. (2011) PLoS ONE, 6, e19791. F€orster, T. (1951) Fluoreszenz Organischer Verbindungen, Vandenhoeck & Ruprecht, G€ottingen. Stryer, L. and Haugland, R.P. (1967) Proceedings of the National Academy of Sciences of the United States of America, 58, 716–726. Vogel, S.S., Thaler, C., Blank, P.S., and Koushik, S.V. (2010) FLIM Microscopy in Biology and Medicine (eds A. Periasamy and R.M. Clegg), Taylor & Francis, Boca Raton, pp. 245–290. Brand, L., Knutson, J.R., Davenport, L., Beechem, J., Dale, R.E. et al. (1985) Spectroscopy and the Dynamics of Biological Systems (eds P.M. Bayley and R.E. Dale), Academic Press, New York, pp. 259–305. Lakowicz, J.R. (2006) Principles of Fluorescence Spectroscopy, Springer, New York, p. 954. Dale, R.E. (1978) Acta Physica Polonica, A54, 743–756. Vogel, S.S., Nguyen, T.A., van der Meer, B.W., and Blank, P.S. (2012) PLoS ONE, 7, e49593. Rizzo, M.A., Springer, G.H., Granada, B., and Piston, D.W. (2004) Nature Biotechnology, 22, 445–449. Nagai, T., Ibata, K., Park, E.S., Kubota, M., and Mikoshiba, K. (2002) Nature Biotechnology, 20, 87–90. Koushik, S.V., Chen, H., Thaler, C., Puhl, H.L., 3rd, and Vogel, S.S. (2006) Biophysical Journal, 91, L99–L101. Thaler, C., Koushik, S.V., Blank, P.S., and Vogel, S.S. (2005) Biophysical Journal, 89, 2736–2749. Koushik, S.V. and Vogel, S.S. (2008) Journal of Biomedical Optics, 13, 031204. Padilla-Parra, S., Auduge, N., Lalucque, H., Mevel, J.C., and Coppey-Moisan, M. (2009) Biophysical Journal, 97, 2368–2376. van der Meer, B.W., Raymer, M.A., Wagoner, S.L., Hackney, R.L., Beechem, J.M., and Gratton, E. (1993) Biophysical Journal, 64, 1243–1263. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 104 5 How to Apply FRET: From Experimental Design to Data Analysis Niko Hildebrandt 5.1 Introduction: FRET – More Than a Four-Letter Word! F€orster resonance energy transfer (FRET) is a very special scientific topic because it inspires and challenges many theoretical and experimental scientists from different research disciplines ranging from the fundamental life sciences over theoretical physics and chemistry up to applied technology in physics, electronics, chemistry, medicine, and biology. Almost as versatile as the topic itself is the discussion about the acronym FRET or rather about the first letter “F.” Should it be “F” like fluorescence, which is most often involved in FRET experiments (although FRET is a nonradiative energy transfer)? Or “F” like F€ orster, who was the first person to develop a theory relating spectroscopic data such as absorption and emission spectra to the energy transfer efficiency, donor–acceptor molecule distances and orientations (although many other scientists were involved in the discovery of FRET)? Or should the “F” be completely erased in order to circumvent the discussion (or to avoid four-letter words)? Personally, I prefer to use “F€ orster” in acknowledgment of his achievements and in order to avoid the term fluorescence, but probably the most important aspect of this discussion is the fact that it is not an important discussion. FRET is a very useful and interesting technology and its experimental application and theoretical treatment for the many possible FRET systems should be investigated, discussed, and developed. Although the main theory was contributed by F€orster in the 1940s, FRET is a very modern technology because the r6 distance dependence over approximately 1–20 nm fits extremely well into the recent discoveries and investigations in nanoscience and nanobiotechnology [1]. The everincreasing number of donor–acceptor pairs (cf. Chapter 14) is another evidence for the contemporary relevance of FRET. This chapter will cover the main aspects of what FRET is, what FRET can be used for (and for what it should better not be used), how a FRET experiment should be designed and performed, what mathematics is absolutely necessary, and how the experimental results can be processed and interpreted. It must also be mentioned that there is no one recipe for all FRET experiments. The main part of this chapter should rather be understood as a guide to FRET, providing useful information that FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j105 j 5 How to Apply FRET: From Experimental Design to Data Analysis must be considered for every FRET experiment in order to receive the correct answers from the investigated systems. Combination with the theory described in Chapters 3 and 4 and the specific tools and applications presented at the end of this chapter as well as the following application chapters will further enlarge the scope of possibilities of the powerful FRET technology. Moreover, there are several excellent monographs and review articles available [2–20] that are related to FRET application for more specific problems that are not covered within this chapter and/or book. 5.2 FRET: Let’s get started! Before treating the concept, some inevitable mathematics, and the experimental work related to FRET, I would like to introduce some first ideas about what FRET can be useful for and what kind of prethoughts should be taken into consideration. If these first questions are clarified, one can continue with understanding FRET in order to perform and interpret a successful experiment. Most of the aspects mentioned in the following paragraph might seem evident and even more than obvious for the FRET expert. However, often such first simple thoughts can avoid later trouble, when the project and/or experiment have already been planned or maybe even carried out. FRET is a strongly distance-dependent transfer of energy between two molecules. This energy transfer will take place only at a distance range of approximately 1–20 nm, and if the system to be studied (the object of interest) does not provide such distances, FRET is not a good solution for its investigation. If the distance range is possibly met by the system, two molecules are required between which FRET can take place – the so-called FRET pair. The energy donor (D) must be a luminophore [luminescent molecule or particle – it should be noted that fluorescence (singlet– singlet optical transition) is a subterm of luminescence, which is the general term for the emission of light originating from the electronic transition between two different energy states [15,17,18]]. The energy acceptor (A) must be able to absorb light in the same spectral range as the emission of D. The absorption and emission spectra of the FRET pair should be chosen (e.g., from Chapter 14) in a suitable wavelength range, which fits to the available instrumentation (excitation source and detection setup) and does not interfere with the object of interest (e.g., excitation or emission of some components within the investigated system). Moreover, it should be taken care that the optical properties of the FRET pair are preserved within the environment of the object of interest (e.g., for experiments that need to be performed in aqueous solution, D and A should be avoided that are only soluble in organic solvents). Once a good FRETpair has been identified, it should be ensured that D and A can be attached to the object of interest (e.g., bioconjugation to a protein) and that the properties of the object as well as of D and A are only minimally influenced by this conjugation. Another aspect to be taken into consideration is if the object of interest will be studied on the single-molecule level or in an ensemble of many objects. A single FRET pair gives only yes or no answers (FRET was successful or not) and many of these FRET pairs (or many excitations of the same FRET pair) need to be analyzed in order to calculate FRET efficiencies. The ensemble Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 106 measurement results in averaged luminescence intensities or lifetimes, which must be analyzed to reveal FRET efficiencies. One should also keep in mind that (due to the additional FRET energy path) the overall luminescence quantum yield of a FRET pair (no matter if the luminescence of D or A is measured) is in the most cases lower than the luminescence quantum yields of D or A alone. Therefore, the use of FRET only makes sense in cases where it can provide specific information on two sites within the object of interest due to D–A proximity (ligand–receptor or antibody– antigen binding, distances of two specific positions within a biological molecule, colocalization of two molecules, ratiometric D–A sensor, reduced photobleaching of A by excitation via D, etc.). If only one event needs to be monitored (e.g., the binding of a specific antibody to the cell membrane), pure luminescence measurements (using excitation and emission without FRET) are usually the preferred method of choice and the additional step of FRET can be avoided. Scheme 5.1 illustrates the main aspects that should be taken into consideration before thinking about an application of FRET. Scheme 5.1 Flowchart for prethoughts concerning the application of FRET. 5.3 FRET: The Basic Concept FRET is an energy transfer process between a luminescent donor molecule or particle (the donor D) and a light-absorbing acceptor molecule or particle (the acceptor A). The luminescence energy (in spectroscopy and imaging usually expressed in wavelength, but wavenumber can also be found) of D must be equal to the absorption energy (wavelength) of A, which is referred to as resonance condition (FRET ¼ F€ orster Resonance EnergyTransfer).TheenergyistransferrednonradiativelyfromDtoAwithanefficiency that is dependent on the distance r between D and A (gFRET r6). The origins of this r6 j107 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.3 FRET: The Basic Concept j 5 How to Apply FRET: From Experimental Design to Data Analysis distance dependence were discovered long before F€ orster’s contributions to FRET and a very nice historic overview by Robert Clegg describes the early findings of energy transfer between two molecules separated by distances beyond orbital overlap [21]. FRET is based on the approximation that dipole–dipole coupling can be represented by Coulombic coupling (coupling of two charges) VCoul. In fact, VCoul should be dominant at the usually considered FRET distance range of approximately 1–20 nm, where orbital overlap-related mechanisms (for very short distances) and radiative mechanisms (for long distances) play minor roles. The FRET rate can be represented by Fermi’s golden rule: kFRET ¼ 2p 2 jVj r; h ð5:1Þ where h is the reduced Planck constant, V is the electronic coupling between D and A, and r is the density of the interacting initial and final energetic states, which is related to the spectral overlap integral J, describing the overlap of D emission and A absorption (see below). In Equation 5.1, V can be replaced by the r3 distance dependent VCoul: V Coul ¼ kj~ mD k ~ mAj ; 4pe0 n2 r 3 ð5:2Þ where ~ mD and ~ mA are the transition dipole moments of D and A, k is the orientation factor between them (cf. Equation 5.9), e0 is the vacuum permittivity, n is the refractive index, and r is the distance between D and A. By substituting Equation 5.2 into Equation 5.1, one arrives at the r6 distance dependence of the FRET rate: kFRET ¼ 9ðln 10Þk2 WD J; 128p5 N A n4 tD r 6 ð5:3Þ where WD is the luminescence quantum yield of D, NA is Avogadro’s number, and tD is the luminescence lifetime of D (in the absence of FRET). Figure 5.1 shows the basic principle of FRET, including the Coulombic mechanism, where an electronic transition from a higher to a lower energy level in D leads to an electronic transition from a lower to a higher energy level in A (without electron exchange!) if these transitions are in energetic resonance. At a distance r, where the FRET rate and all other decay rates are in equilibrium 1 ðkFRET ¼ kRD þ kNR D ¼ tD Þ, the FRET efficiency gFRET is 50%. This distance is the socalled F€orster distance (or F€orster radius) R0, which can be calculated by replacing kFRET with t1 D and r with R0 in Equation 5.3: 1=6 9ðln 10Þk2 WD R0 ¼ J : 128p5 N A n4 ð5:4Þ J is the spectral overlap integral [defined in the wavelength (l) or wavenumber (~ n) scale]: ð ð ~ n J ¼ I D ðlÞeA ðlÞl4 dl ¼ I D ð~nÞeA ð~ nÞd 4 ; ð5:5Þ ~ n Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 108 Figure 5.1 Basic FRET principle. (a) Simplified energy level scheme (Jablonski diagram) representing the excitation of the donor (hn) from an electronic ground state (D) to an excited state (D ), followed by inner relaxation (vibrational and rotational – dotted arrow) to an excited electronic ground state, followed by radiative decay (kR), nonradiative decay (kNR), or FRET (kFRET). In order to realize the FRET process from D to A , the difference between the respective energy levels need to be equal (resonance condition: E(D ) E(D) ¼ E(A ) E (A)), as emphasized by the coupled transitions (horizontal lines with dots on each end). After FRET, the acceptor is in an excited state (A ), followed by radiative or nonradiative decay to its ground state (A). (b) Different energy pathways after donor excitation (hnex) possibly leading to luminescence emission of D (hnD) or A (hnA) for FRET analysis. which is dependent on the acceptor molar absorptivity (or extinction coefficient) spectrum eA and the donor emission spectrum ID normalized to unity (cf. Figure 5.2): ð ð I D ðlÞdl ¼ I D ð~nÞd~n ¼ 1: ð5:6Þ Combination of Equations 5.3 and 5.4 leads to the relation between the FRET rate, the luminescence decay time of the donor, and the distances (r6 distance Figure 5.2 The overlap (gray area) of the areanormalized emission spectrum of D (cf. Equation 5.6) and the extinction coefficient (or molar absorptivity) spectrum of A (eA) defines the overlap integral J (cf. Equation 5.5), which is directly proportional to the FRET rate (cf. Equation 5.3). In this graph, a wavelength scale was chosen (wavenumber is also possible – cf. Equations 5.5 and 5.6). j109 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.3 FRET: The Basic Concept j 5 How to Apply FRET: From Experimental Design to Data Analysis 1.0 ηFRET 0.8 0.6 0.4 0.2 0.0 0.5R0 R0 1.5R0 2R0 2.5R0 r Figure 5.3 FRET efficiency (gFRET) as a function of D–A distance (r). The r6 distance dependence (cf. Equation 5.8a) leads to a strong sensitivity of gFRET to r in the distance region of about 0.5R0–2.0R0 (gray background area). dependence of the FRET rate): 6 1 R0 kFRET ¼ tD : r ð5:7Þ The FRET efficiency is then gFRET ¼ kFRET 1 ¼ : kFRET þ t1 1 þ ðr=R0 Þ6 D ð5:8aÞ As shown in Figure 5.3, the sensitivity of the FRET efficiency to the D–A distance is most prominent in a region between about 0.5R0 and 2.0R0, with the efficiency curve being very steep around R0 (high dynamic range). The last important variable for the basic FRET concept is the so-called orientation factor kappa-squared (k2 in Equations 5.3 and 5.4 or k in Equation 5.2), which is often forgotten to be taken into account seriously and is also often considered too much in detail even though a relatively easy averaging might be applicable. In this regard, it is very important if the FRET experiment is performed on a single-molecule level (where only averaging over time will make sense) or the ensemble (where averaging over time and/or over the molecular ensemble might be reasonable). In any case, one should seriously think about kappa-squared for any FRET application and then choose the most appropriate option of treating the dipole–dipole orientation. Figure 5.4 describes the orientation of the donor and acceptor dipoles within the basic concept of FRET. Taking the different angles between the transition dipole moments of D and A (~ mD and ~ mA ) and the connection vector between D and A (~ r ) allows the calculation of the orientation factor: ^D m ^A 3ð^ k¼m mD ^r Þð^ mA ^r Þ ¼ cos qDA 3cos qD cos qA ; ð5:9Þ Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 110 Figure 5.4 Orientation of the donor emission transition dipole moment ~ mD , the acceptor r for the calculation of absorption transition dipole moment ~ mA , and the D–A connection vector~ the FRET dipole orientation factor k (Equation 5.9). ^D , m ^A , and ^r represent the unit vectors of ~ where m mD , ~ mA , and~ r , respectively, and the angles qDA, qD, and qA are shown in Figure 5.4. There are some reasonable averaging conditions, which can provide a good approximation for k2 to be used for many practical FRET applications. When all D’s and A’s can take any possible orientation during the FRET time (1/kFRET), which means that the average rotation rate is much larger than the average FRET rate ðkrot kFRET Þ, the system is in a dynamic averaging regime and k2 becomes 2/3. This fast isotropic rotation of D and A is often fulfilled if they are bound to polypeptides or proteins [20]. If D and A are both luminescent, fast isotropic rotation can be verified by checking if their emissions are unpolarized. If one of the FRET partners shows average orientation (isotropically degenerate: “sphere”) and the other has a fixed orientation (well-defined linear dipole: “line”), then k2 (from “sphere” to “line” or from “line” to “sphere”) can take values between 1/3 and 4/3 (for which 2/3 is still not such a bad approximation). In case of FRET from “line” to “line” (two well-defined linear dipoles), it becomes much more complicated because the full orientation factor range (0 < k2 <4) needs to be considered (and 2/3 might be a very bad approximation). In the case where all donors and acceptors are fixed (no rotational motion), each FRET pair is assumed to be isolated from all other pairs, and the electronic transitions of D and A are single dipoles, one can use a static regime approximation [11] for which k2 is dependent on the D–A distance r and the D–A distance r0, for which k2 would be 2/3. In this case, k2 can take values between 0 (for r 0.4r0) and 2/3 (for r 1.4r0). As already mentioned, it is important to take into account the orientation of the D and A transition dipole moments for every FRET experiment in order to justify an approximation or the assumption of a special orientation. If the experiments allow a modification of the D–A distance (so that FRET can be measured at multiple different distances), one can try to evaluate the FRET efficiency results using F€ orster distances (R0) calculated with different k2 values in order to get a better idea about which orientational approximation or assumption might lead to reasonable results. In any case, an estimated or exact k2 can be calculated and a detailed treatment of the orientation factor, including the presentation of how to access k2 values, can be found in Chapter 4. j111 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.3 FRET: The Basic Concept j 5 How to Apply FRET: From Experimental Design to Data Analysis 5.4 FRET: Inevitable Mathematics Understanding the basic concept of FRET and the r6 distance dependence requires the equations from Section 5.3. For a profound knowledge of FRET theory and an exact theoretical calculation of different FRET parameters, Chapters 3 and 4 as well as the available literature should be consulted [3,4,10,11,15,17–20,22–27]. As already mentioned in Section 5.1, many different research disciplines make use of FRET and especially in applied research, it is sometimes tried to avoid mathematics if possible. However, even the FRET experimentalist will need some mathematics for the choice of the adequate D–A pair and for the interpretation of the results. In most cases, only very few (and relatively) uncomplicated equations are necessary in order to achieve very satisfactory experimental results. The most important FRET parameters are the F€orster distance R0 and the FRET efficiency gFRET because they combine spectroscopic data (e.g., luminescence quantum yields, intensities, and lifetimes or emission and absorption spectra) with distances and orientations. 5.4.1 F€ orster Distance (or F€orster Radius) Equation 5.4 is a general equation that needs a careful choice of units for achieving the correct value and unit for the resulting distances and/or FRET rates. For the experimental case, one can predefine commonly used units within the overlap integral J and merge all constants found in Equations 5.4 into one value: 9ðln 10Þ ¼ 8:79 128p5 N A 1028 mol; ð5:10Þ where Avogadro’s constant NA ¼ 6.02 214 1023 mol1. Different examples of facilitating Equation 5.4 with preassuming different units can be found in Chapter 3. Probably the most common length unit used at the small distance scale of FRET is nanometer and, therefore, it makes sense to express r and R0 in nanometers. Quite often the unit Angstr€om is also used for FRET (1 A ¼ 0.1 nm). Moreover, simplification of Equation 5.4 can be achieved by using commonly used units for optical spectroscopy in the overlap integral (Equation 5.5): the wavelength l in nanometer units and the molar absorptivity (or extinction coefficient) eA in M1 cm1 (liter per mol per centimeter). These predefinitions lead to the following F€ orster distance (for which k2, WD, n4, and J are dimensionless): R0 ¼ ð8:79 1028 Taking into account that M k2 WD n4 J mol M1 cm1 nm4 Þ1=6 : 1 1 cm ð5:11Þ 1 nm ¼ 10 nm mol , this leads to R0 ¼ 0:02108ðk2 WD n4 JÞ1=6 nm; 4 17 6 ð5:12Þ which is the F€orster distance in nanometers. It is very important that Equations 5.11 and 5.12 are valid only if the value for J is calculated in M1 cm1 nm4. Using different units will lead to a different prefactor on the right-hand side of Equation 5.12. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 112 5.4.2 FRET Efficiency The FRET efficiency relates the F€orster distance (R0, which can be calculated by absorption and emission spectroscopy of D and A as demonstrated above) and the D–A distance (r, which might be known from the system to be measured or which might be the unknown variable of interest) with spectroscopic data of D and A as single entities and as a D–A FRET pair. The FRET efficiency gFRET can be determined by several different methods, which are explained in the following. 5.4.2.1 Determination by Donor Quenching One possibility of calculating the FRET efficiency is to use spectroscopic data (luminescence quantum yield W, lifetime t, and intensity I) of D in the absence (subscript D) or presence of A (subscript DA). Using Equation 5.8b, kFRET kFRET ¼ ¼ kFRET tDA kFRET þ t1 kFRET þ kRD þ kNR D D gFRET ¼ ð5:8bÞ R and taking into account that WD ¼ tD kRD ¼ kRD =ðkRD þ kNR D Þ and WDA ¼ tDA kD ¼ kRD =ðkFRET þ kRD þ kNR D Þ, this leads to the following equation: 1 gFRET ¼ 1 þ ðr=R0 Þ 6 ¼ R60 R60 WDA tDA I DA ¼1 ¼1 ¼1 : WD tD ID þ r6 ð5:13Þ The last part of Equation 5.13 (using emission intensities) is valid only if the excitation light intensity absorbed by D and all the measurement parameters are identical for both measurements (D in the absence and in the presence of A). If the experimental conditions for “D” and “DA” FRET measurements are similar, this is usually a good approximation leading to adequate results. FRET causes quenching of the donor luminescence quantum yield, lifetime, and intensity and thus WDA, tDA, and IDA are smaller than WD, tD, and ID, leading to efficiency values between 1 and 0. This technique requires the determination of WD, tD, or ID before the FRET measurement and gFRET is then calculated from data generated by the two different experiments. Once R0 has been calculated (Equation 5.12), the spectroscopic data can be used for calculating the D–A distance r by converting Equation 5.13 to r ¼ R0 WDA WD WDA 1=6 ¼ R0 tDA tD tDA 1=6 ¼ R0 I DA ID I DA 1=6 : ð5:14Þ 5.4.2.2 Determination by Acceptor Sensitization Donor quenching is not an evidence of FRET, as this donor deactivation can also be caused by other quenching mechanisms. The only sure evidence of energy transfer from D to A is to measure the luminescence of A after excitation of D. Obviously, a luminescent acceptor is necessary for this, so this technique cannot be performed with dark quenchers as FRET acceptors. The FRET efficiency can then be calculated by the ratio of acceptor luminescence intensity in the presence (IAD) and in the j113 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.4 FRET: Inevitable Mathematics j 5 How to Apply FRET: From Experimental Design to Data Analysis absence (IA) of D. For the calculation of gFRET by sensitized acceptor luminescence, one needs to take into account that there is (in almost all cases) direct excitation of A at the excitation wavelength used for exciting D. Although this direct excitation might be weak, it needs to be corrected for in order to achieve appropriate results for gFRET. This correction can be done by subtracting the direct luminescence IA from IAD and multiplying the corrected luminescence ratio by the ratio of the absorptivities (or extinction coefficients) of A and D (eA and eD) at the excitation wavelength used for the FRET experiment. The FRET efficiency is then 1 R60 I AD I A eA IAD eA gFRET ¼ ¼ ¼ 1 ¼ : IA eD IA eD 1 þ ðr=R0 Þ6 R60 þ r 6 ð5:15Þ This technique requires the knowledge of eA and eD and the measurement of IA before the FRET experiment. gFRET is then calculated from data generated by the two different measurements. Once R0 has been calculated (Equation 5.12), the spectroscopic data can be used for calculating the D–A distance r by converting Equation 5.15 to 1=6 I A eD r ¼ R0 1 : ð5:16Þ ðI AD I A ÞeA Cases of incomplete D–A labeling (free A inside the FRET sample) and the use of two different excitation wavelengths (for cases of weak FRET where the ratio IAD/IA is close to unity, which can cause significant errors in calculating gFRET) using this technique with further necessary corrections have been applied by Clegg et al. [28–30]. 5.4.2.3 Determination by Donor Quenching and Acceptor Sensitization In order to calculate gFRET from a single measurement, the quenched donor luminescence intensity (D in the presence of A) and the sensitized acceptor luminescence intensity (A in the presence of D) can be analyzed. This technique can offer the advantage of high precision in calculating the FRET efficiency because data from simultaneously measured luminescence spectra are used. However, the acceptor luminescence caused by direct excitation of A still needs to be corrected for and thus it needs to be measured by a preexperiment. Moreover, the luminescence quantum yields of D and A need to be known. For the calculation of gFRET, the excitations (intensity divided by quantum yield: I/W) of D and A are taken into account. In this case, the FRET efficiency can be defined as the number of donor excitations leading to acceptor excitations (FRET) divided by all donor excitations (leading to FRET and all other radiative and nonradiative deactivation pathways): R60 ðI AD I A Þ=WA ¼ 6 6 6 ðI IA Þ=WA þ I DA =WD R þ r 1 þ ðr=R0 Þ AD 0 1 ðI AD IA Þ WA I DA ¼ ¼ 1þ : WD I AD IA ðI AD IA Þ þ ðWA =WD ÞI DA gFRET ¼ 1 ¼ ð5:17aÞ Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 114 In this equation, IAD is the luminescence intensity of A during FRET (in the presence of D), which needs to be corrected for acceptor luminescence due to direct excitation (IA). IDA is the luminescence intensity of D during FRET (in the presence of A) and WA and WD are the luminescence quantum yields of A and D, respectively. Once R0 has been calculated (Equation 5.12), the spectroscopic data can be used for calculating the D–A distance r by converting Equation 5.17a to 1=6 WA IDA r ¼ R0 : ð5:18Þ WD ðI AD I A Þ 5.4.2.4 Determination by Donor Photobleaching Apart from measuring the luminescence quenching of donor and/or acceptor, another possibility to determine the FRET efficiency is photobleaching, which is especially useful in imaging setups, where the necessary light sources, providing enough power for photobleaching, are often readily available. Using the photobleaching time constant tbl of D in the absence (subscript D) or in the presence of A (subscript DA), gFRET can be calculated [16,31]: 1 gFRET ¼ 1 þ ðr=R0 Þ6 ¼ R60 tbl D ¼ 1 : R60 þ r 6 tbl DA ð5:19Þ In contrast to the luminescence properties W, t and I in Equation 5.13, the photobleaching time constant of D in the absence of A (tbl D , no FRET) is found in the numerator, whereas the photobleaching time constant of D in the presence of A (tbl DA , FRET) is found in the denominator on the right-hand side of Equation 5.19. bl In this case, tbl D is smaller than tDA because FRET opens a new energy pathway for the excited donor to return to the ground state and, therefore, the photobleaching time constant of the donor is increased in the presence of the acceptor [16]. Similar to the D or A quenching approaches, this technique requires the determination of tbl D before the FRET measurement and gFRET is then calculated from data generated by the two different experiments (no FRET and FRET). Once R0 has been calculated (Equation 5.12), the D–A distance r can be calculated by converting Equation 5.19 to r ¼ R0 tbl D bl tDA tbl D 1=6 : ð5:20Þ 5.4.2.5 Determination by Acceptor Photobleaching In order to be able to determine gFRET from a single sample, one can use acceptor photobleaching. In this approach, the initially FRET-quenched donor luminescence (D in the presence of A) is recovered by photobleaching the acceptor (destroying the FRETpath to A). The FRETefficiency can then be calculated using the luminescence intensities of D before (superscript pre) and after (superscript post) the photobleaching of A: gFRET ¼ 1 1 þ ðr=R0 Þ6 pre ¼ R60 I DA ¼ 1 post : 6 6 R0 þ r I DA ð5:21Þ j115 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.4 FRET: Inevitable Mathematics j 5 How to Apply FRET: From Experimental Design to Data Analysis Note that donor and acceptor are still physically connected after photobleaching, but the acceptor (and the FRET path) is irreversibly “switched off.” The main advantage of this technique is the avoidance of measuring two different samples (such as pure D and photobleached D in the previous paragraph). This can be especially problematic for cellular imaging, for example, due to different protein expression levels from cell to cell, which will cause variable donor concentrations in the different cells. Once R0 has been calculated (Equation 5.12), the photobleaching data can be used for calculating the D–A distance r by converting Equation 5.21 to pre r ¼ R0 I DA post pre I DA IDA !1=6 : ð5:22Þ 5.4.3 FRET with Multiple Donors and/or Acceptors In many FRETexperiments, the interaction between multiple donors and acceptors is possible. For example, this can be the case for random labeling of proteins, cell surfaces, or nanoparticles, where the amount of D and A that can possibly interact is mainly defined by the density of D and A within the labeled systems. There can be different or equal distances between the D’s and A’s, there can be very few D’s interacting with A’s and vice versa, there can be D’s and A’s that do not interact at all, and there can be one-, two-, and three-dimensional distributions that can be random or contain excluded spaces (so that random distribution cannot be assumed anymore). Raicu has proposed a theoretical model for multiple donor–acceptor interactions, which becomes more complicated if more possibilities are included in the model [32]. Using an approximation of equal distances between all D’s and A’s and assuming that all FRETrates are equal for any D–A FRETpair, Raicu derived an equation for the FRET efficiency (gmulti FRET ), which is purely dependent on the number of acceptors nA (and not on the number of donors) and the efficiency of a single D–A pair (gFRET): gmulti FRET ¼ nA gFRET : 1 þ ðnA 1ÞgFRET ð5:23Þ The same result was found by Clapp et al. in an experimental study for single D to multiple A FRET, where several organic dye acceptors were placed around one quasispherical semiconductor quantum dot donor in order to achieve approximately equal distances [33]. The authors derived (and confirmed experimentally) the following equation: gmulti FRET ¼ nA kFRET nA R60 ¼ ; 1 nA kFRET þ tD nA R60 þ r 6 ð5:24Þ which is equal to Equation 5.23. An increasing FRET efficiency with an increasing number of A’s per D can be explained by the fact that there are simply more (nA instead of 1 for a single pair) possible de-excitation pathways available for the excited donor and, therefore, the probability of de-exciting D via FRET becomes higher. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 116 The independence of the FRET efficiency from the number of donors is not so obvious, as in the case of multiple donors with one acceptor, the multiple D’s need to compete for FRET to the single A (unless there is always only one D or A excited during the time of excitation, energy transfer, and de-excitation). This problem has been taken into account by Berney and Danuser (with further developments by Corry et al. for different geometrical distributions) [34,35]. As the integration of the many possible parameters for multiple D–A systems inside an analytical model can become very complicated, numerical approaches are an excellent alternative. The authors developed a Monte Carlo simulation (MCS) for an experimental setup in which they could control donor and acceptor concentrations (and thereby the D to A ratio RDA) as well as the average distance between D and A using a PLL-g-PEG-biotin-coated microscope coverslip surface and controlled amounts of streptavidin labeled with D or A. The MCS results were used as a reference to compare different methods from the literature to use the experimental data for FRET efficiency determination. The beauty of the numerical approach is the stepwise calculation of the FRETefficiency photon by photon (or exciton by exciton). In the following, the MCS scheme is briefly outlined [34,35]: Step 1: The coordinates and types of D’s and A’s are assigned, accounting for a regular arrangement and for excluded volume effects. Step 2: The transfer probability from each donor Di to every acceptor Aj is calculated by P ij ¼ R60 =r 6ij ; ð5:25Þ where R0 is the F€orster distance and rij is the distance between the Di–Aj FRET pairs. Step 3: The exciton flux (dependent on the photon flux and the extinction coefficient and concentration of the donors) is calculated by 2 ð5:26Þ f e ¼ pa2 Ilh1 c1 1 10½ðeD nD Þ=ð1000N A pa Þ ; where a is the radius of the simulated system, I is the irradiance of the excitation light source with wavelength l, h is the Planck’s constant (6.626 1034 Js), c is the speed of light (3 108 ms1), eD is the molar extinction coefficient of the donor at wavelength l, nD is the number of donor molecules, and NA is the Avogadro’s constant (6.022 1023 mol1). Step 4: A time sequence giving the play time of each exciton within the time interval of excitons being incident on the fluorophores is defined; for each exciton, a target donor is randomly assigned and the experimental clock is set to zero. Step 5: The excitons are played (sequentially) to see if they are absorbed by the donor and, if yes, whether it is de-excited by FRET or fluorescence. First it is checked if the donor is already “busy” (already excited) and if yes the exciton is lost and the next one will be played. If the donor is not “busy”, it gets excited, is then set to “busy” and a list of free acceptors (which are not already excited) around the donor is generated. The overall rate of energy release is calculated by ! afree X 1 1 ð5:27Þ tT ¼ tD 1 þ Pij ; j¼1 j117 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.4 FRET: Inevitable Mathematics j 5 How to Apply FRET: From Experimental Design to Data Analysis where tD is the luminescence lifetime of D in the absence of A (unquenched). The time for the donor to release its energy in the simulation is calculated by T D ¼ tT ln ðcD Þ; ð5:28Þ where cD is a uniformly distributed random generator delivering a value between 0 and 1. TD defines the time point of energy release of the excited D (the time point at which the donor is set from “busy” back to “free”). Next it needs to be decided (using probabilities) whether Di will release its energy as FRET or fluorescence. This is determined by creating a cumulative histogram with the classes of all possible pathways, where the probability of the classes FRET to Aj is (tT/tD)Pij and the probability of the class fluorescence is tT/tD. Another uniform random number between 0 and 1 is picked to decide for the de-excitation class. If the selection falls in the class of fluorescence, the variable Fluo is incremented by 1 and the next exciton is played. If Di was selected for FRET to Aj, this acceptor is set to “busy” and the variable FRET is incremented by 1. The time interval of a “busy” Aj (TA) is determined by a MCS step similar to Equation 5.28, where TA ¼ tA ln (cA) with tA denoting the luminescence lifetime of A and cA is another random number between 0 and 1. The complete step 5 is repeated for all excitons. Step 6: Finally, the simulated FRET efficiency can be calculated by comparing the number of donors undergoing FRET and fluorescence, respectively: gFRET ¼ FRET : FRET þ Fluo ð5:29Þ 5.5 FRET: The Experiment This section will cover the most important aspects of how to design, perform, and analyze a general FRET experiment using steady-state and time-resolved optical spectroscopy and microscopy and applying the equations from the previous section. 5.5.1 The Donor–Acceptor FRET Pair A well-designed FRET experiment can provide a lot of useful qualitative and quantitative information. As already mentioned in Section 5.2, every FRET experiment starts with considering how the system of interest can be most efficiently analyzed by FRET, which is closely related to the choice of a suitable donor–acceptor FRET pair. One of the first considerations should concern the physical parameter to be determined, which could be the following: 1) Quantitative distance: FRET as a spectroscopic (or molecular) ruler to determine the distance between two molecules in a range of about 1–20 nm. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 118 2) Qualitative distance: FRET to prove a proximity (colocalization) between two molecules or FRET as an optical switch between two molecules in about 1–20 nm distance. 3) Quantitative concentration: FRET to determine the amount of two (or more) bound molecules in about 1–20 nm distance. 4) Qualitative (or semiquantitative) concentration: FRET to prove the existence of two (or more) bound molecules in about 1–20 nm distance. In all cases, a convenient FRET pair should be chosen, which combines bright luminescence of D and/or A for an efficient detection or signal generation with a sufficiently large R0 for the distance between the two molecules. It is not always advisable to choose the highest possible F€orster distance. For example, if R0 ¼ 6 nm and the D–A distance is r ¼ 2 nm, the FRET efficiency will be very close to 100% (WDA ¼ IDA ¼ tDA ¼ 0) leading to a pure on–off luminescence signal, which could cause problems in evaluating and/or quantifying the physical phenomenon behind (disruption of D–A binding and disappearance of D, photobleaching of D, FRET due to very close proximity, and other quenching mechanisms independent of FRET). In most cases, the FRET pair should be chosen such that the distance (or distances) of interest is in the 0.5R0–2.0R0 distance range (cf. Figure 5.3) because this is the most sensitive distance range for FRET. Another important aspect is the technique with which FRETwill be analyzed (donor quenching and/or acceptor sensitization, donor or acceptor photobleaching) (cf. Section 5.4) as well as the available equipment because this defines the photophysical properties (absorption and emission wavelength range, quantum yield, lifetime, etc.) of D and A. Many fluorophores are available for FRET, such as organic dyes (and dark quenchers), polymeric and dendrimeric dyes, naturally occurring fluorophores, lanthanide, and metal-based complexes, as well as nano- and microparticles. For a detailed overview of the different fluorophores, the reader is referred to Chapter 6 and to Ref. [1], in which the important topic of bioconjugation (how to attach D and A to the system of interest) is also outlined. The most comprehensive Chapter 14 concerning FRET data contains much information about F€ orster distances, photophysical properties, availability, and applications of many D–A FRETpairs. It is always recommended to choose at least three fluorophores, that is, one preferred D–A FRET pair with at least one alternative D and/or A for control experiments or backup solution in case the original FRET pair does not perform sufficiently well. 5.5.2 F€ orster Distance Determination After having made the initial choice of a D–A FRET pair, the F€ orster distance should be determined (or verified in case an R0 value could already be found in the literature). Assuming that sufficient donor and acceptor material is available (especially absorption measurements require higher concentrations than fluorescence detection) and that the orientation factor can be sufficiently well estimated or calculated, the determination of R0 is usually not a very challenging task, which j119 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis requires standard steady-state spectroscopy equipment such as absorption and fluorescence spectrometers. Both spectra (donor emission and acceptor absorption) should be measured as precisely as possible, but in a reasonable manner. Some suggestions of acquiring such spectra are outlined below. More details can be found, for example, in Refs [15,18]. a) Molar absorptivity (or extinction coefficient) spectrum of the acceptor: As many absorption spectrometers (especially plate readers) provide lower spectral resolution (often limited to minimum 1 nm) than fluorescence spectrometers (usually well below 1 nm), it is recommended to start with measuring the absorption spectrum of A (usually the absorbance A or optical density OD is measured for a light path length of 1 cm). For the calculation of R0 in a spreadsheet program, it is quite convenient to use the same wavelength steps for emission and absorption spectra. One can save time erasing out data points on the computer by already recording the spectra with matching wavelengths. It does not make much sense to measure the emission spectrum of D (ID(l)) with a 10-fold higher spectral resolution than the absorption spectrum of A because both spectra are multiplied in the overlap integral (Equation 5.5). However, care must be taken when spectra with very narrow peaks (e.g., lanthanide emission spectra) are used, for which the spectral resolution should be high enough not to omit emission (or absorption) peaks by not measuring them because of too large wavelength steps or by erasing them in a spreadsheet when calculating the overlap integral. Apart from the acceptor sample, the pure solvent (or buffer) should be measured and subtracted from the sample spectrum (background correction). The OD (for 1 cm light path) of the sample should not be too low because this can cause large errors after background correction. On the other hand, too high ODs can cause saturation. In most cases, a maximum absorption intensity between 0.1 and 3 OD should provide satisfactory results. Measuring an emission excitation spectrum (variation of the excitation wavelength while recording a fixed emission wavelength) of the same but highly diluted acceptor sample on a fluorescence spectrometer is a good option to qualitatively (shape of the spectrum) verify the absorbance spectrum. Once the background-corrected absorbance spectrum A(l) has been determined, it can be calculated into the molar extinction coefficient spectrum e(l) using Lambert–Beer’s law: AðlÞ ¼ eðlÞ c l ) eðlÞ ¼ AðlÞ ; cl ð5:30Þ where c is the concentration of the sample and l is the light path length (e.g., 1 cm for a standard cuvette). b) Emission spectrum of the donor: Most emission spectra (unless they have very narrow emission bands) can be measured quite precisely by recording an emission intensity every 0.5 or 1.0 nm (wavelength step). As already mentioned, it is recommended to use the same wavelength steps as for the absorption spectra if possible. Important aspects for recording emission spectra are to avoid high sample concentrations (which can cause inner filter effects), solvent and/or cuvette contaminations (which might contain other luminescent materials) and Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 120 0.008 ID(λ) ε A(λ) 1.50x105 1.25x105 1.00x105 0.006 7.50x104 0.004 5.00x104 0.002 0.000 400 2.50x104 450 500 550 wavelength (nm) extinction coefficient (M–1 cm–1) normalized emission intensity M–1cm–1 0.00 600 Figure 5.5 Area-normalized (cf. Equation 5.6) donor emission spectrum (left – cf. column 3 in Table 5.1), acceptor molar extinction coefficient spectrum (right – cf. column 4 in Table 5.1), and resulting overlap function (shaded spectrum in the center – cf. column 5 in Table 5.1). light scattering (Raman peaks inside the emission spectrum), and to use an adequate quantum correction (detector sensitivity changes as a function of wavelength) in order to obtain a correctly shaped emission spectrum. As the spectrum is area-normalized for the calculation of R0, total intensities are not important. However, the relative intensities must be correct. As an example of calculating R0, I chose two imaginary D and A molecules with emission and absorption spectra showing the typical shape of organic dye spectra with a pronounced maximum and a blueshifted (for absorption) or redshifted (for emission) “shoulder.” The spectra cover a wavelength range of about 400–600 nm and overlap in the 450–550 nm region (Figure 5.5). Once the spectra are recorded, the overlap integral and F€ orster distance can be easily calculated using a spreadsheet program (e.g., Excel or Origin), as shown in Table 5.1. The first column contains the wavelengths from 400–600 nm (in the presented case, 0.5 nm steps are used) and the second one contains the determined emission spectrum ID(l) (in arbitrary units – e.g., photon counts). The sum of all ID(l) values is calculated (last cell of the second column) and used to calculate the third column by dividing each ID(l) value with this sum in order to get the areanormalized emission spectrum I D ðlÞ (as a control, the sum of column 3 should be unity). The fourth column contains the extinction coefficient spectrum of A (eA(l) in M1 cm1 units) and the fifth column is the product of l4 I D ðlÞ eA ðlÞ. The sum of this column 5 is the overlap integral J, which can be used (together with the predetermined values for k2, WD, and n) to calculate R0 (in nm) using Equation 5.12 (as shown in the lower part of Table 5.1 with some arbitrarily chosen values for k2, WD, and n). j121 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis Table 5.1 Extract from a spreadsheet to calculate R0 from donor emission and acceptor absorption spectra (cf. Figure 5.5) and predetermined values of k2, WD, and n. l (nm) ID(l) (arb. units) ID(l) eA(l) (M – 1 cm – 1) l4ID(l) eA(l) (nm4 M – 1 cm – 1) 400 400.5 401 401.5 402 402.5 403 .. . 478.5 479 479.5 480 480.5 481 481.5 .. . 513.5 514 514.5 515 515.5 516 516.5 .. . 597.0 597.5 598.0 598.5 599.0 599.5 600.0 Sum 0 0 0 0 0 0 0 .. . 3.10E þ 05 3.12E þ 05 3.13E þ 05 3.14E þ 05 3.14E þ 05 3.14E þ 05 3.14E þ 05 .. . 1.20E þ 05 1.19E þ 05 1.18E þ 05 1.17E þ 05 1.16E þ 05 1.15E þ 05 1.14E þ 05 .. . 0 0 0 0 0 0 0 3.3E þ 07 0 0 0 0 0 0 0 .. . 9.52E 03 9.56E 03 9.60E 03 9.62E 03 9.64E 03 9.64E 03 9.64E 03 .. . 3.67E 03 3.64E 03 3.61E 03 3.58E 03 3.55E 03 3.52E 03 3.50E 03 .. . 0 0 0 0 0 0 0 1.0 109.4 119.3 130.1 141.8 154.4 168.1 182.8 .. . 5.35E þ 04 5.36E þ 04 5.38E þ 04 5.40E þ 04 5.42E þ 04 5.45E þ 04 5.48E þ 04 .. . 1.54E þ 05 1.54E þ 05 1.54E þ 05 1.54E þ 05 1.54E þ 05 1.53E þ 05 1.53E þ 05 .. . 0 0 0 0 0 0 0 1.6E þ 07 0 0 0 0 0 0 0 .. . 2.67E þ 13 2.70E þ 13 2.73E þ 13 2.76E þ 13 2.79E þ 13 2.81E þ 13 2.84E þ 13 .. . 3.93E þ 13 3.91E þ 13 3.90E þ 13 3.88E þ 13 3.85E þ 13 3.83E þ 13 3.80E þ 13 .. . 0 0 0 0 0 0 0 4.6E þ 15 J (nm4 M – 1 cm – 1) k2 WD n R0 (nm) 4.6E þ 15 0.67 0.55 1.40 5.8 5.5.3 The Main FRET Experiment With the photophysical properties and the F€ orster distance of the D–A FRET pair in the pocket, the real FRET experiment can begin. This section will focus on luminescence spectroscopy. FRET microscopy techniques have been recently treated in two Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 122 comprehensive books [4,16]. Photobleaching FRET is mainly used for imaging and has been covered in Ref. [36], so it will not be treated again in this section. The main FRET experiment will always include luminescence, because the emission changes of D or A or both must be measured in order to determine FRET efficiencies and/or distances. Luminescence quenching can occur due to many different reasons and therefore it is always recommended to measure acceptor sensitization in combination with donor quenching. As the sensitization of A must occur from energy transfer (assuming that direct excitation of A has been taken into account and properly subtracted), the acceptor detection channel can be coined as the “FRET-proof channel.” The best solution is to compare D-quenching and Asensitization and if the FRET efficiencies can be calculated as equal for both partners, the involved mechanism is most probably FRET (or another energy or charge transfer process – cf. Section 5.6). Luminescence can be measured using steady-state and/or time-resolved techniques. The latter can be divided into timedomain and frequency-domain techniques. Details about the different experimental approaches (steady-state and time-resolved) can be found in Refs [15,18], and only a short technical overview will be presented here. In order to draw the correct conclusions from the experimental data, it is important to combine as many experiments as possible (D-quenching, A-sensitization, steady-state and timeresolved measurements, and all the necessary control experiments). The aim should be the same as it used to be for Theodor F€orster: to search for the most appropriate solution of a scientific problem, or in his favorite term “The correct interpretation of an observation” (Die richtige Deutung einer Beobachtung) (cf. Chapter 1). 5.5.3.1 Steady-State FRET Measurements Steady-state luminescence spectroscopy measures emission spectra, that is, the emission intensity of the luminophores as a function of wavelength (or wavenumber) over their complete time period of emission (no temporal resolution). This is usually achieved by exciting the luminophores at a fixed wavelength, while scanning (e.g., with a monochromator) over the wavelength of their emission, which is then detected [e.g., by using a photomultiplier tube (PMT)] in selected wavelength intervals. Measuring the full emission spectra offers the advantage that both D and A can be measured simultaneously within the same sample. The values of ID, IDA, IA, and IAD can then be extracted from the measured overall spectrum of the sample (containing both D and A emissions) and used with Equations 5.13–5.18 for the calculation of FRETparameters and distances. A correct treatment of the overall spectrum requires a deconvolution of the two spectra followed by integration over each single spectrum in order to obtain correct values for ID, IDA, IA, and IAD (cf. Figure 5.6). Depending on how spectrally close the two emission spectra are, the spectral cross talk between them will be weaker (for well-separated spectra) or stronger (for close spectra). In case of well-separated emission spectra, it might be sufficient to simply take the peak intensity values of D and A to obtain ID, IDA, IA, or IAD. Due to the almost endless choice of possible FRET pairs (which strongly depend on the application), a generic example of steady-state FRETresults with an imaginary FRET pair (the same as already chosen in Figure 5.5 for the calculation of the F€ orster j123 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis (a) (c) Figure 5.6 Absorption and emission spectra resulting from a representative steady-state FRET experiment with luminescent D and A (cf. Figure 5.5 for calculation of R0). (a) Peaknormalized absorption (dotted) and emission spectra of D in the absence of A (black) and A in the absence of D (gray). (b) Emission spectra of different mixtures of D and A excited at 420 nm. FRET causes quenching of donor luminescence and sensitization of acceptor luminescence. Therefore, the spectra are composed of different ratios of the emission spectra of D and A [from part (a)]. (c and d) Deconvolution of the D–A spectra leads to the emission spectra of D (b) (d) in the presence of A [part (c)] and of A in the presence of D [part (d)]. The black curves in both graphs represent the emission spectra of D and A in the absence of A and D, respectively (donor emission without FRET quenching and acceptor emission due to direct excitation at 420 nm without FRET sensitization). Integrating over these emission spectra results in the intensity values ID and IA. The gray curves represent the quenched donor and sensitized acceptor emission, respectively. Integrating over these emission spectra results in the intensity values IDA and (IAD þ IA) (cf. Equations 5.13–5.18). distance R0) is presented in Figure 5.6. The principle of this theoretical example can be transferred to any practical experiment using real FRET pairs. Many practical examples using different FRET pairs can be found in Chapters 6–14 and in the FRET literature. Figure 5.6 presents the spectroscopic data obtained from the main FRET experiment. Part (a) shows the emission (and absorption) spectra of pure D and pure A, which need to be measured before the FRET experiment, but under the same experimental conditions (i.e., concentration, solvent, excitation and emission conditions, etc.). Part (b) presents the luminescence spectra obtained from 11 different Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 124 j125 FRET measurements (11 different samples or 1 sample with changing spectra over time), for which donor quenching and concomitant acceptor sensitization become quite obvious. In the case where no third luminescent components (e.g., sample autofluorescence) are present within these spectra, they are composed of different ratios of the single emission spectra from part (a). Thus, they can be deconvoluted to single FRET spectra of D and A, as shown in parts (c) and (d). Integration over these spectra results in the intensities ID, IDA, IA, and IAD, which are necessary to perform the FRET calculations using Equations 5.13–5.18. With the obtained intensity values, different interpretations concerning FRET are possible. We will begin with the analysis of the donor spectra and evaluate afterward, which advantages the acceptor spectra provide for the interpretation of the FRET system. 1) One can analyze the results in Figure 5.6c using donor quenching (Section 5.4.2.1). As summarized in Table 5.2, the normalized donor emission intensity IDA/ID decreases from 1 to 0 (in steps of 0.1), which can be caused by several reasons: a) Distance: Assuming that each D is connected with one A (complete labeling), the FRET efficiencies gFRET of each spectrum can be calculated by Equation 5.13 (Table 5.2, column 4). Using the precalculated R0 of 5.8 nm (Figure 5.5 and Table 5.1), the D–A distances r can be calculated by Equation 5.14 (Table 5.2, column 5). This leads to the conclusion that all different Table 5.2 Data resulting from different interpretations of D-quenching (Figure 5.6c) and A-sensitization (Figure 5.6d) using steady-state FRET measurements. 0 1 2 Experimental data Measurement 1 2 3 4 5 6 7 8 9 10 11 3 4a) 5b) Distance 6c) 7d) Concentration quenching IDA/ID IAD/IA (IAD – IA)/ID gFRET r (nm) KSV [Q] [DA]/[D]0 or [DA]/[DA]max 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 1.00 2.43 3.86 5.29 6.71 8.14 9.57 11.00 12.43 13.86 15.29 0.000 0.054 0.108 0.162 0.216 0.270 0.324 0.378 0.432 0.486 0.540 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 >13.0 8.4 7.3 6.7 6.2 5.8 5.4 5.0 4.6 4.0 0.0 0.00 0.11 0.25 0.43 0.67 1.00 1.50 2.33 4.00 9.00 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 8e) Multiple A’s nA with r ¼ 6.7 nm 0 Not measured Not measured 1.0 1.6 2.3 3.5 5.4 9.3 21.0 Not measured The different results are visualized in Figure 5.7. a) gFRET calculated with Equation 5.13 for donor quenching, Equation 5.15 for acceptor sensitization using eA/ eD ¼ 0.07, and Equation 5.17a for combined donor quenching and acceptor sensitization using WA/WD ¼ 0.54. b) r calculated with Equation 5.14 for donor quenching, Equation 5.16 for acceptor sensitization using eD/eA ¼ 1/ 0.07, and Equation 5.18 for combined donor quenching and acceptor sensitization using WA/WD ¼ 0.54. c) Cannot be analyzed by acceptor sensitization. d) [DA]/[D]0 ¼ 1 (IDA/ID) for D-quenching and [DA]/[DA]max ¼ (IAD/IA) 1)/(IAD/IA)max 1) for A-sensitization analysis. e) nA calculated with Equation 5.24. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis (b) (c) (d) SV (a) r = 6.7 nm = constant Figure 5.7 Different interpretations of one and the same experiment resulting from the analysis of the donor-quenching emission spectra in Figure 5.6c taken from the calculated data from Table 5.2. (a) Different FRET efficiencies due to different D–A distances. (b) Static or dynamic quenching with increasing quencher concentration. (c) Increasing amount of D–A FRET pairs at fixed D–A distance. (d) Increasing amount of A’s per D at fixed D–A distance. samples contain the same concentration of completely labeled single D–A pairs with different D–A distances (e.g., change of structural confirmation over time or due to changes in the environment, such as temperature or pH), which decrease from measurement 1 to 11, as depicted in Figure 5.7a. b) Concentration quenching: The D-quenching could be caused by any deactivation process (including FRET), which can be described by the Stern–Volmer equation: ID ¼ 1 þ K SV ½Q ; I DA ð5:31Þ where KSV is the Stern–Volmer constant, which is K SV ¼ K S ¼ ½DQ ½D ½Q ð5:32Þ in the case of static quenching and K SV ¼ K D ¼ kq tD ð5:33Þ Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 126 in the case of dynamic quenching. In these equations, [Q], [D], and [DQ] are the respective concentrations of the quencher, the donor, and the donor– quencher complex, kq is the dynamic (or bimolecular) quenching constant, and tD is the luminescence lifetime of D (in the absence of Q). Static or dynamic quenching will lead to a linear quenching behavior (which is not the case for combined static and dynamic quenching) as found for the experiment from Figure 5.6 (and shown in Figure 5.7b and Table 5.2, column 6). In the case of FRET quenching, a possible scenario for which the Stern–Volmer equation can be applied would be a FRET assay, for which increasing amounts of an analyte (e.g., a biomarker) are titrated to a constant concentration of Dand A-labeled specific binding molecules (e.g., antibodies against the biomarker). This case will cause static quenching (because the D–A distance r is fixed) due to addition of the analyte with concentration [Q]. As D and A are in excess, each addition of an analyte will lead to the formation of a D–A pair with the concentration [DA] ¼ [DQ] ¼ [Q] (assuming 100% binding efficiency). With [D] ¼ [D]0 [DA], where [D]0 is the initial concentration of D, Equation 5.32 will become KSV ¼ KS ¼ ([D]0 [DA])1 and Equation 5.31 will become [DA]/[D]0 ¼ 1 (IDA/ID), which is presented in Figure 5.7c and Table 5.2, column 7. c) Multiple acceptors: In the case of experimental results without the spectra from measurements 2, 3, and 11, another possible interpretation from Figure 5.6c would be an increasing amount of A’s per D (nA from Equation 5.24) with a fixed distance r (Figure 5.7d and Table 5.2, column 8). In the case for which r is unknown, Equation 5.24 can be used with r as a free fit parameter (R0 and gFRET are known from the measurements of Figure 5.5 and 5.6c) in order to determine the D–A distance. d) Environmental quenching: Cases (b) and (c) assume stable environmental conditions. If this is not the case and the environment (e.g., temperature, solvent, and pH) changes during the experiment (over time or from sample to sample), luminescencequenching[includingFRETduetodistancechangesasmentioned in case (a)] might also be caused by these factors. Often such changes also lead to deviations in the luminescence and/or absorption spectra (bathochromic or redshift, hypsochromic or blueshift, or intensity changes of different luminescence bands from the same luminescent species). Thus, it is wise to also have a careful look at the spectral features of the different measurements. 2) In addition to donor quenching, one can analyze the results in Figure 5.6d using acceptor sensitization (Section 5.4.2.2). Analyzing the acceptor spectra has the advantage that they can provide compelling evidence for FRET (or at least energy transfer) because they show that A was excited via D, whereas donor quenching could have other reasons than deactivation via the acceptor. As summarized in Table 5.2, IAD/IA increases from 1 to 15.29, which can be explained by the same reasons as for donor quenching and leads to the same results using different equations for their calculation (cf. footnotes in Table 5.2). However, one should keep in mind that the present example is an idealized theoretical FRET experiment and for most “real-world” experiments, a verification of the donor j127 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis IDA (IAD-IA) wavelength (nm) Figure 5.8 (a) FRET quenching of D (decreasing IDA) and FRET sensitization of A (increasing IAD–IA, spectra without emission from direct acceptor excitation) deconvoluted from the spectra of Figure 5.6b (D in the absence of A: black curve). (b) The negative (b) -slope = ΦA/ΦD (IAD-IA) / ID emission intensity (counts) (a) normalized emission intensity quenching results by acceptor sensitization is necessary to draw reliable conclusions about FRET. In the case where D-quenching and A-sensitization provide similar results using the FRET equations, other quenching effects can be excluded. This is highly important for justifying results that are based on FRET, for example, if distances are reported or analyte concentrations are calculated by assuming that all donor quenching is caused by FRET. In our example, the distances calculated for donor quenching can be confirmed by finding similar results from acceptor sensitization analysis (Table 5.2, column 5). Also, concentration quenching (Table 5.2, column 7) can be confirmed by finding similar results from A-sensitization analysis using the D–A concentration ([DA]max) found for the maximum acceptor-sensitized emission intensity instead of the initial donor concentration [D]0. Assuming a fixed distance r, the multiple acceptors can also be confirmed by acceptor sensitization. 3) Instead of analyzing donor quenching and acceptor sensitization separately, they can also be combined within Equations 5.17a and 5.18 for calculating FRET parameters and distances (donor quenching and acceptor sensitization – Section 5.4.2.3). In this case, it is not necessary to know the emission intensity of the pure donor (ID), but the intensity arising from direct excitation of A (IA) must be subtracted from IAD and the luminescence quantum yields of D and A need to be known (cf. Equation 5.17a). As summarized in Table 5.2, IDA decreases, while (IAD IA) increases (columns 1 and 3, both normalized to ID), as already discussed above for donor quenching and acceptor sensitization analysis. Figure 5.8a shows the deconvoluted (from Figure 5.6b) emission intensity spectra of simultaneous donor quenching and acceptor sensitization. Again, the same results are found using different equations for their calculation (cf. footnotes in Table 5.2). Changing Equation 5.17a can be used to calculate luminescence quantum yields of D and A. IDA/ID = 1-ηFRET slope of FRET-sensitized acceptor emission intensity as a function of FRET-quenched donor emission intensity (both intensities normalized to ID) gives the luminescence quantum yield ratio of A and D. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 128 gFRET ¼ I AD I A WA ) gFRET I DA ¼ ðI AD I A Þð1 gFRET Þ ðI AD IA Þ þ ðWA =WD ÞI DA WD ) gFRET WA I DA WA I AD I A WA WA IDA IAD I A I DA ¼ ðI AD I A Þ ) gFRET ¼ ) ¼ : WD ID WD ID WD WD ID ID ð5:17bÞ The negative slope or the intersection with the ordinate of (IAD IA)/ID as a function of IDA/ID (Figure 5.8b) is the luminescence quantum yield ratio of acceptor and donor (WA/WD), and therefore the calculation of unknown D or A quantum yields becomes possible, in case one of them is known. Although all these different possibilities of FRET data analysis exist, many studies use only donor intensity quenching in order to relate the experimental data to FRET. This is often not correct, as many other processes (cf. Section 5.6) can be the cause of donor quenching. The choice of donor quenching, acceptor sensitization, or the combination of both strongly depends on the experimental conditions and the required experimental accuracy. In general, it is more precise to use all the different possibilities and to carefully deconvolve the different emission spectra in order to avoid spectral overlap problems. However, for well-separated spectra, high emission intensities, and many data points supporting the experimental interpretation, it can be sufficient to use only donor quenching. Moreover, highly sensitive luminescence detection applications make use of optical bandpass filters for spectral separation, which means that the full spectral information is not available. In such cases, one should ensure that D and A bandpass filters with minimal spectral overlap to A and D are used or that the emission spectra are measured before the filterbased FRET measurement, such that this information can be combined with the filter transmission spectra to achieve an adequate spectral correction. Measuring emission spectra without an adequate quantum correction (correcting the wavelength-dependent detection efficiency of photodetectors such as PMTs), autofluorescence from the sample medium, sample scattering and reabsorption (inner filter effects), and low signal-to-noise ratios are common sources of error that should be taken into account for FRET measurements. Please keep in mind that the theoretical generic example presented here was chosen to demonstrate the different possibilities of interpreting a FRET experiment. Therefore, donor quenching, acceptor sensitization, and their combination deliver exactly the same results for FRET calculations. As reality is usually not so kind to adapt all energy transfer or luminescence quenching systems to FRET theory, the different approaches leading to different results will usually be quite helpful to find reasonable explanations for the experimental data. As the large variety of different energetic excitation and relaxation processes involved in FRETexperiments can lead to many paths of energy flow and therefore many paths of interpretation, one should always consider a reasonable amount of control experiments adapted to each individual FRET system of interest. j129 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis 5.5.3.2 Time-Resolved FRET Measurements Time-resolved luminescence spectroscopy can be divided into two technologies for measuring luminescence decay times, namely, time-domain and frequency-domain measurements [15,18]. Time-domain methods measure the time-dependent luminescence intensity of a sample following a short excitation pulse of light (usually in the nanosecond range for lasers and up to some microseconds for flash lamps – preferably, the excitation pulse is much shorter than the decay time of the sample). Frequencydomain methods use intensity-modulated (usually sinusoidal modulation) excitation light (e.g., I ¼ Iav(ex) þ Ip(ex) cos vt). The modulation frequency ( f ¼ v/2p) is typically in the same range as the reciprocal of the luminescence decay time of the sample (e.g., 100 MHz ¼ 1/(10 ns)). The emitted light will follow this modulation frequency, but with a time delay (e.g., I ¼ Iav(em) þ Ip(em) cos (vt w)), which is usually called phase shift or phase angle (w). Moreover, the peak intensity (Ip(ex) and Ip(em) for excitation and emission, respectively) will be lower and the average intensity (Iav(ex) and Iav(em)) can be different. This is usually expressed in the modulation ratio M ¼ [Ip(em)/Iav(em)]/[Ip(ex)/ Iav(ex)]. Both methods (time-domain and frequency-domain) (cf. Figure 5.9) can be used to determine single or multiple luminescence decay times (ti). The mathematical description of a time-dependent luminescence decay (with i decay times) is the luminescence intensity (I) as a function of time: X X t t ¼A : ð5:34Þ I¼ Ai exp ai exp ti ti i i For the frequency domain, the phase shift (w) or the modulation ratio (M) can be used for the determination of single or multiple decay times. The mathematical relation between phase shift and decay times is P ðai vt2i Þ=ð1 þ v2 t2i Þ w ¼ arctan Pi : ð5:35Þ 2 2 i ðai ti Þ=ð1 þ v ti Þ (b) 1.0 0.8 0.75 intensity intensity (a) 1.00 0.50 0.25 Ip(em) Ip(ex) 0.6 0.4 Iav(em) Iav(ex) 0.2 φ 0.00 0 2 4 6 time t (ns) 8 10 Figure 5.9 Examples of time-domain (a) and frequency-domain (b) measurements for the determination of luminescence decay times. Excitation is displayed in black [(a): usually many pulses, for example, with 80 MHz repetition rate or one pulse every 12.5 ns, are 12 0.0 0 10 20 30 time t (ns) 40 50 necessary to record a full decay curve; (b): modulated intensity with 80 MHz modulation frequency). Emission with a decay time of 2 ns is displayed in gray [(a): intensity decay; (b): modulated intensity with 80 MHz). Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 130 (b) 1 phase shift φ (º) or modulation ratio M (%) luminescence intensity I (a) 0.1 0 5 10 15 time t (ns) 20 Figure 5.10 (a) Luminescence intensity decay curve resulting from a time-domain measurement. (b) Phase shift (gray) and modulation ratio (black) curves resulting from frequency-domain measurement. For both 100 80 60 40 20 0 0.1 25 1 10 100 1000 10 000 frequency f (MHz) results, the decay times are t1 ¼ 2 ns (with a1 ¼ 0.7) and t2 ¼ 14 ns (with a2 ¼ 0.3). The single lifetime components are shown in the dash-dotted (2 ns) and dotted (14 ns) curves. The mathematical relation between modulation ratio and decay times is "P M¼ 2 i ðai vti Þ=ð1 #1=2 P 2 2 2 2 þ v2 t2i Þ þ i ðai ti Þ=ð1 þ v ti Þ : P 2 i ai ti ð5:36Þ Equations 5.34–5.36 can be used within a least-square analysis, for which the parameters ai and ti are varied until a best fit between the experimental data and the mathematical fit values is achieved. Figure 5.10 shows typical curves for timedomain and frequency-domain data, which were generated for a double-exponential luminescence decay using Equations 5.34–5.36. No matter how the luminescence decay times for a FRET system are measured, they can be used to determine FRET efficiencies and distances (in case R0 was determined before). In most cases, donor quenching (Equations 5.13 and 5.14) is used for decay time FRET analysis (for decay time analysis using photobleaching, refer to Ref. [36]). FRET analysis of sensitized acceptor luminescence decays is usually complicated because both the donor and the acceptor excited states are involved in FRET. Therefore, the acceptor decay will be a combination of the FRETquenched donor excited-state lifetime (tDA) and the acceptor excited-state lifetime (tA) [37]. The concentration of an excited acceptor ([A ]) after excitation of the donor can be expressed as a differential equation: d½A ¼ kFRET D ðkRA þ kNR A Þ A ; dt j131 ð5:37Þ where [D ] is the excited donor concentration and kRA and kNR A are the radiative and nonradiative decay rates of the acceptor. In this equation, kFRET ½D describes the increase in the A excited-state population due to FRET from excited D and ðkRA þ kNR represents radiative and nonradiative deactivation of excited A. A Þ½A Solving Equation 5.37 and assuming the donor decay as single-exponential lead Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis to an excited-state acceptor concentration as a function of time (luminescence decay function of excited-state acceptors) [18]: ½D kFRET t ½D kFRET t 1 0 1 exp A ¼ 1 0 1 þ ½A 0 exp tA tDA tDA tA tDA tA ð5:38aÞ or A ¼ ½D 0 gFRET þ ½A 1 ðtDA =tA Þ 0 ½D 0 gFRET t t exp ; exp 1 ðtDA =tA Þ tA tDA ð5:38bÞ where [D ]0 and [A ]0 are the initial (at t ¼ 0) concentrations of excited D and A, respectively, and tA is the luminescence decay time of A in the absence of D. As already mentioned, these acceptor excited-state decay functions (representing the decay of A in the presence of D with decay time tAD) are composed of the “pure” acceptor decay (first term in Equation 5.38a or 5.38b) and the FRET-quenched donor decay (increase of the acceptor luminescence with the time-component tDA represented by the negative term in Equation 5.38a or 5.38b). Figure 5.11 shows the excited-state decay curves (with decay time tAD) of FRET-sensitized A (in the presence of D) for different tA values in comparison to decay curves of FRET-quenched D (in the presence of A with decay time tDA) and pure D (in the absence of A with decay time tD). These curves were calculated from Equation 5.38b for gFRET values of 50 and 95%, respectively, and assuming no direct acceptor excitation ([A ]0 ¼ 0). Figure 5.11 shows that the decay times tAD (or the slopes of the black curves) and tDA (or the slope of the gray curves) are equal for tA tD (gray dotted curves for tD). The higher the FRET efficiency, the larger the required difference between tA and tD (for the case of 95% efficiency, Figure 5.11b, a value of tA ¼ 0.1tD already shows (a) Figure 5.11 Excited-state decay curves of A with tA values of 0.001, 0.01, 0.1, 0.5, 1, 2, and 5 times tD, respectively – black curves from bottom to top calculated with Equation 5.38b for FRET efficiencies of gFRET ¼ 50% (a) and (b) gFRET ¼ 95% (b). Pure donor decays with decay time tD (dotted gray curves) and FRETquenched donor decays with decay times tDA (gray curves) are shown for comparison. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 132 clear differences in tAD and tDA). Figure 5.11 also visualizes that the FRET decay curves for similar tA and tD values can become complicated as they show a rise time at the beginning of the acceptor decay curves and significantly longer decays than for tDA. The use of donors with long decay times (e.g., lanthanide complexes as treated in Section 5.6) and acceptors with short decay times (e.g., organic dyes) allows the replacement of tDA by tAD in Equations 5.13 and 5.14 and, therefore, a time-resolved FRET analysis of quenched donor emission and sensitized acceptor emission. One big advantage of the analysis within the “FRET-proof” acceptor channel is the absence of pure donor emission (D in the absence of A) and therefore a lower background signal (non-FRET signal). 5.5.3.3 Interpretation of Time-Resolved FRET Data In this section, we will only treat donor quenching (acceptor-sensitized timeresolved analysis will be treated in the next section) and only decay time results from time-domain measurements will be shown (frequency-domain measurements will result in the same luminescence decay times and therefore lead to the same FRET results). In the example of Figure 5.6, the luminescence intensity of the donor is quenched from 100 to 0% of the initial (D in the absence of A) donor intensity (in steps of 10%). Assuming a monoexponential luminescence decay function (I ¼ A exp(t/t) with an arbitrary chosen luminescence decay time of t ¼ 5 ns) of the unquenched donor (D in the absence of A), there are several possible scenarios of time-dependant donor (D in the presence of A) luminescence intensity decays, which can lead to the steady-state spectra presented in Figure 5.6. The amplitudes (AD and ADA) and decay times (tD and tDA) of the luminescence decays can provide useful information about the investigated FRET system concerning static (concentration-dependent) and dynamic (distance-dependent) quenching or a mixture of both. This information cannot be found by analyzing only the steady-state spectra. A comparison of steady-state and time-resolved quenching in the so-called Stern–Volmer plots (Figure 5.12) can give good evidence of the quenching situation. In the case where the ratio of initial to quenched steady-state intensity (I0/I) and the ratio of initial to quenched decay time (t0/t) increase equally over quencher (a) (b) (c) I0 /I higher temperature 1 [Q] τ0 /τ τ0 /τ τ0 /τ = I0/I 1 τ0 /τ or I0/I I0 /I τ0 /τ or I0/I τ0 /τ or I0/I higher temperature 1 [Q] [Q] Figure 5.12 Stern–Vomler plots of dynamic (a), static (b), and combined dynamic and static (c) quenching. The influence of temperature on dynamic and static quenching is also indicated within the graphs. j133 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis (b) normalized luminescence intensity (a) (c) (d) measurement Figure 5.13 (a and b) [part (b) intensity normalized with logarithmic intensity scale] Dynamic luminescence quenching represented by decay curves (D in the presence of A – gray) with constant amplitudes and decreasing decay times compared to the unquenched D curve (black). (c) Unchanged amplitudes (circles) and decreasing decay times (squares) as a function of intensity quenching. (d) The purely dynamic quenching behavior is confirmed by a Stern– Volmer plot, for which both the intensity ratio (triangles) and the decay time ratio (squares) increase equally from measurement to measurement. concentration [Q], the quenching is purely dynamic. If the intensity ratio increases over quencher concentration and the decay time ratio is unaffected, the quenching is purely static. If both the intensity ratio and the decay time ratio increase over quencher concentration but the intensity ratio increases stronger, the quenching is a mixture of dynamic and static deactivation. Different from the steady-state results, where one set of spectra (experimental data in Table 5.2) was used to provide different possible interpretations, we will discuss here different sets of possible decay curves, which can confirm or disprove the different possible interpretations from the steady-state measurements. Thus, the different time-resolved scenarios (Figures 5.13–5.16 and Tables 5.3–5.5) will be discussed in relation to the steady-state experimental data (spectra from Figure 5.6). 1) Distance (dynamic quenching) As already mentioned in Section 5.5.3.1, increased quenching of D can be explained by decreased D–A distances. If the measured samples lead to the spectra Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 134 (b) (c) (d) normalized luminescence intensity (a) measurement Figure 5.14 (a and b) [part (b) intensity normalized with logarithmic intensity scale] Static luminescence quenching represented by decay curves (D in the presence of A – gray) with constant decay times and decreasing amplitudes compared to the unquenched D curve (black). (c) Decreasing amplitudes (circles) and unchanged decay times (squares) as a function of intensity quenching. (d) The purely static quenching behavior is confirmed by a Stern–Volmer plot, for which the intensity ratio (triangles) increases, whereas the decay time ratio (squares) is unchanged. in Figure 5.6c (steady-state measurements) in combination with the decay curves from Figure 5.13, the FRET quenching was most probably caused by decreasing D–A distances (e.g., change of structural confirmation over time or due to changes in the environment, such as temperature or pH), as depicted in Figure 5.7a. Within the decay time functions (Equation 5.34), the amplitudes ADA (amplitudes for D in the presence of A) stay at a constant value (ADA/AD ¼ 1), whereas the decay times tDA decrease (tDA/tD decreases from 1 to 0 in steps of 0.1), which is well illustrated by the decreasing slopes in the logarithmic plot. The Stern–Volmer plot (for which the quencher concentration was replaced by the measurement number – in case concentrations are known, they can be used instead) shows an equal increase of intensity and decay time ratio, which confirms the purely dynamic quenching behavior. The FRET efficiencies gFRET (Equation 5.13) and D–A distances (Equation 5.14) calculated by using luminescence intensities (Table 5.2, columns 4 and 5) are confirmed by the values calculated by using luminescence decay times (Table 5.3, columns 4 and 5). j135 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis (b) normalized luminescence intensity (a) (c) (d) measurement Figure 5.15 (a and b) [part (b) intensity normalized with logarithmic intensity scale] Combined dynamic and static luminescence quenching represented by decay curves (D in the presence of A – gray) with changing decay times and amplitudes compared to the unquenched D curve (black). (c) Changing (overall decreasing tendency) amplitudes (circles) and changing (overall decreasing tendency) decay times (squares) as a function of intensity quenching. (d) The combined dynamic and static quenching behavior is confirmed by a Stern–Volmer plot, for which the intensity ratio (triangles) increases stronger than the decay time ratio (squares). 2) Concentration or environmental conditions (static quenching) If the steady-state spectra from Figure 5.6c are accompanied by the decay curves from Figure 5.14, the FRET quenching was most probably caused by increasing concentrations of D–A pairs described by Equation 5.32 (with A being the quencher Q) or a change in environmental conditions (e.g., temperature, solvent, and pH). Within the decay curves, the amplitudes ADA decrease (ADA/AD decreases from 1 to 0 in steps of 0.1), whereas the decay times tDA stay at a constant value (tDA/tD ¼ 1), which becomes quite obvious in the logarithmic plot with all decay curves showing the same slope. The Stern–Volmer plot shows an increase of intensity ratio in combination with a constant decay time ratio, which confirms the purely static quenching behavior. In the case of FRET quenching, a possible scenario would be a FRET assay, for which increasing amounts of an analyte (e.g., a biomarker) are titrated to a constant concentration of D- and Alabeled specific binding molecules (e.g., antibodies against the biomarker). Although FRET is a dynamic quenching process, this case will cause static Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 136 Table 5.3 Data resulting from purely dynamic or purely static D-quenching (Figures 5.13 and 5.14). 0 1 2 3 4a) 5b) 6 Measurement Overall quenching Distance (dynamic quenching) IDA/ID 1 2 3 4 5 6 7 8 9 10 11 1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 7 8c) Concentration (static quenching) t DA/t D ADA/AD gFRET r (nm) t DA/t D ADA/AD [DA]/[D]0 1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 >13.0 8.4 7.3 6.7 6.2 5.8 5.4 5.0 4.6 4.0 0.0 a) gFRET calculated with Equation 5.13. b) r calculated with Equation 5.14. c) [DA]/[D] 0 ¼ 1 (I DA/I D) ¼ 1 (A DA tDA/A D tD) ¼ 1 (A DA/A D ) measurements. 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 1.0 because 1.0 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 tDA/tD ¼ 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 for all Table 5.4 Data resulting from a combination of dynamic and static D-quenching (Figure 5.15). 0 1 Measurement 1 2 3 4 5 6 7 8 9 10 11 3a) 2 4 5b) Dynamic and static quenching t DA/tD gFRET r (nm) ADA/AD [DA]/[D]0 1.0 0.989 0.952 0.921 0.968 0.685 0.784 0.476 0.357 0.256 0 0.0 0.011 0.048 0.079 0.032 0.315 0.216 0.524 0.643 0.744 1.0 >13.0 12 9.6 8.7 10 6.6 7.2 5.7 5.3 4.9 0 1.0 0.910 0.840 0.760 0.620 0.730 0.510 0.630 0.560 0.390 0 0.000 0.090 0.160 0.240 0.380 0.270 0.490 0.370 0.440 0.610 1.000 a) r calculated with Equation 5.14. b) [DA]/[D]0 ¼ 1 (ADA/AD). j137 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.5 FRET: The Experiment j 5 How to Apply FRET: From Experimental Design to Data Analysis quenching (because the D–A distance r is fixed). The D–A pair concentration ([DA]/[D0]) increases in steps of 10% (Table 5.3, column 8). 3) Multiple acceptors (dynamic quenching) The steady-state spectra from Figure 5.6c and the decay times from Figure 5.13 (in case measurements 2, 3, and 11 are not taken into account) can also be caused by an increasing amount of A’s per D (nA from Equation 5.24) with a fixed distance r (Figure 5.7d and Table 5.2, column 8). Although the distance between D and the multiple A’s is fixed, this configuration leads to dynamic quenching because the FRET efficiency increases with an increasing number of A’s per D. This is not the case for static quenching of an increasing number of single D–A pairs (same FRET efficiency), as discussed in the previous paragraph. In the case for which r is unknown, Equation 5.24 can be used with r as a free fit parameter (R0 and gFRET are known from the measurements of Figure 5.5 and 5.6c) in order to determine the D–A distance. A very nice study of multiple dye acceptors per semiconductor quantum dots was performed by Clapp et al.. The authors used steady-state donor quenching, steady-state acceptor sensitization, and timeresolved donor quenching to find the correct interpretation of their FRET system [33]. 4) Distance and concentration (dynamic and static quenching) If the steady-state and time-resolved measurements lead to the spectra from Figure 5.6c and the decay curves from Figure 5.15, the luminescence deactivation was caused by a combination of dynamic (e.g., decreasing D–A distances) and static quenching (e.g., increasing concentrations of D–A pairs or a change in environmental conditions). Within the decay curves, the amplitudes ADA as well as the decay times tDA change. This change usually has a decreasing tendency for both ADA and tDA (luminescence is quenched), but as amplitude and decay time can compensate for each other, an interpretation of the different decay curves, amplitudes, and decay times might not be as facile as for the pure dynamic or static quenching cases. The Stern–Volmer plot shows a stronger increase of intensity ratio compared to the decay time ratio, which confirms the combination of dynamic and static quenching. In the case of FRET quenching, one of the most important aspects of the time-resolved measurements is the possibility of calculating distances (using the decay times and Equation 5.14). This would not be possible for steady-state measurements because the dynamic (decay times) and the static (amplitudes) quenching parts cannot be distinguished. 5) Multiexponential donor decays and multiple distances FRET systems can be much more complicated than within the simulated examples mentioned above. The first complication can already be the pure donor (D in the absence of A) luminescence decay because it must not necessarily be monoexponential. In an ensemble measurement, the multiple donor molecules can take different conformations (e.g., structural or chemical configurations), which might lead to different decay times for each species and thus an overall multiexponential luminescence decay function. This means that each of these multiple decay times (tDi) or an average decay time (ktDi) must be taken as the pure donor decay time. Averaging is performed using Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 138 amplitude-averaged decay times, because FRET is a dynamic quenching process and a detected signal at a particular time interval is proportional to the excited-state population (and not to the integrated intensity, for which the intensity-averaged decay time would be used) [13,38]. Taking the multiexponential luminescence decay function from Equation 5.34 leads to the following amplitude-averaged decay time: P A i ti X hti ¼ P ¼ ai ti : ð5:39Þ Ai Another aspect making FRET analysis more complicated is the possibility of having multiple D–A distances within a D–A pair ensemble. This means that (apart from the multiexponential tD values) the quenched donor decay times (tDAi) will be multiexponential and depending on the number of different distances (and thus decay times), the time-resolved analysis can become complicated and high signal-to-noise ratios are necessary to recover the different amplitudes and decay times. As if these two difficulties were not enough, complicated FRET systems can also be composed of dynamic and static quenching processes, which means that a careful distinction of amplitudes and decay times (which can compensate for each other within a least-square fit) is necessary for a correct interpretation of the experimental data. In Section 5.6, we will discuss such a FRET system with multiexponential donor decay, multiple distances, and mixed dynamic and static quenching. Moreover, this system uses semiconductor quantum dots as acceptors, which poses another problem, because these nanoparticles are excited at any wavelength below their emission wavelength. Steady-state measurements with such acceptors are useless in most cases because the acceptor will be excited very efficiently (more efficiently than the donor) and the interpretation of quenching and sensitization becomes very difficult. However, the use of lanthanide-based donors with very long excited-state lifetimes (up to several milliseconds) allows the time-resolved analysis of acceptor sensitization due to the large difference in donor and acceptor excited-state lifetimes (cf. Figure 5.11). 5.6 FRET beyond F€ orster In more than six decades that have passed since F€ orster’s paper “Energiewanderung und Fluoreszenz” [22] from 1946, the FRET toolbox has been filled up with many new possible donor and acceptor fluorophores, including a multiplicity of organic dyes, metal-based chelates, fluorescent proteins, and nanoparticles (cf. Chapters 6 and 14). Within all the FRET applications that have been developed, one can find many “nonclassical” approaches, such as FRET from lanthanide donors with multiplet–multiplet transitions (quintet–septet in the case of Eu and Tb) and multiple transition dipole moments, FRET with quantum dot nanoparticles (with diameters of up to 10 nm), which do not present an ideal point-dipole system, and j139 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.6 FRET beyond F€orster j 5 How to Apply FRET: From Experimental Design to Data Analysis FRET without the use of donor excitation by light [bioluminescence resonance energy transfer (BRET) and chemiluminescence resonance energy transfer (CRET)] or energy transfer to metal nanoparticles, which has been related to FRET, nanosurface energy transfer (NSET), dipole to metal particle energy transfer (DMPET), or nanoparticle-induced lifetime modification (NPILM). Moreover, there are other energy transfer mechanisms, such as plasmonic coupling, charge transfer, or singlet oxygen transfer, which can enlarge the distance range of two interacting species from about 1 to 200 nm. In this section we will discuss such systems, which go beyond the classical treatment of Theodor F€ orster. 5.6.1 Time-Resolved FRET with Lanthanide-Based Donors The many advantages of lanthanide-based donors have been known for more than two decades, and especially Paul Selvin should be mentioned here for a lot of pioneering work in this area [9,39–45]. Probably the most important property of lanthanide-based donors for FRET is their long luminescence decay time reaching up to several milliseconds for some supramolecular lanthanide complexes (e.g., chelates or cryptates) [46–52]. This means that the excited-state lifetimes of most lanthanide-based donors are several orders of magnitude larger than those of any other acceptor. Thus, the same decay time analysis can be applied for D-quenching and A-sensitization (cf. Figure 5.11). Moreover, the sensitized acceptor emission can be measured against a very low background. This can be achieved by using an acceptor that emits at a wavelength region void of lanthanide emission (cf. Figure 5.16 for Tb as donor) for minimizing the background of the donor. Many different acceptors (e.g., fluorescent proteins, organic dyes, and quantum dots) are available for the best choice of emission wavelength [1,53,54]. In order to suppress the acceptor background (directly excited acceptor emission), which is usually in the nanosecond time range, one can use pulsed excitation and gate the detector off for a Figure 5.16 Typical emission spectrum of a Tb chelate. The arrows indicate wavelength areas in which an acceptor can be measured without Tb background emission. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 140 short period of time (i.e., several microseconds). This will lead to a pure FRET signal after the detector has been gated on again because almost any photon that will be detected then must arise from an excitation via the donor – the acceptor detection channel becomes a “FRET-proof” channel. This also means that the sensitized acceptor signal is insensitive to concentration effects and incomplete labeling and binding because only those species containing both D and A can contribute to the FRET signal (pure D and pure A signals are completely suppressed). Another advantage is the possibility of large overlap integrals with different acceptors due to the multiple emission bands of lanthanides over a wide spectral range. F€orster distances of 9 nm for an Eu chelate donor and an APC acceptor [55] and up to 11 nm for a Tb chelate donor and quantum dot acceptors [53] have been demonstrated. Such large R0 values can significantly increase the FRET distance range up to about 20 nm (or larger in case of very sensitive detection). One very comfortable aspect of Eu- and Tb-based (most often used lanthanides for FRET applications) donors is their unpolarized emission. Due to their multiple transition dipole moments, they act as randomized donors (even in the absence of fast isotropic rotation) and the orientation factor k2 is limited to values between 1/3 and 4/3 even if the acceptor has a fixed orientation (cf. averaging conditions for k2 in Section 5.3). The following example (lanthanide to quantum dot FRET) will illustrate all the different aspects (including donor quenching, acceptor sensitization, and dynamic and static quenching) of a sophisticated time-resolved FRET analysis using lanthanide-based donors. 5.6.1.1 Terbium to Quantum Dot FRET Using Time-Resolved Donor Quenching and Acceptor Sensitization Analysis In a recent example, we performed a time-resolved analysis of one Tb chelate donor and different quantum dot acceptors in a FRET system, for which Tb and QD are brought in proximity via biotin–streptavidin binding [56]. In this configuration, several Tb donors can attach to the surface of QDs (consisting of the central semiconductor QD and a polymer-based coating for biocompatibility). The luminescence decay of the Tb donor is double-exponential with an amplitude-averaged decay time of ktDi ¼ 2.3 ms and a luminescence quantum yield of 67%. The QD (we will only discuss one QD here) has an emission wavelength maximum of 655 nm, a multiexponential decay in the nanosecond time range (about 30 ns average decay time), and a luminescence quantum yield of 7%. Due to the random labeling of the Tb donor to the streptavidin protein and the ellipsoidal shape of the QD, the FRET system consists of a D–A distance distribution. The different samples contain a constant concentration of Tb donors (0.2 nM) and increasing concentrations of QD acceptors (0 0.6 nM). Due to the varying ratios of Tb/QD in the different samples, different concentrations of pure Tb donors and QD acceptors are present in the FRET systems. The pure Tb donors (at different concentrations) result in varying intensities (amplitudes) of Tb background emission at a fixed decay time (ktDi), which leads to different static contributions with significant intensities in the donor detection channel (optical bandpass filter: 494 20 nm) and minor intensities in the acceptor detection channel (optical bandpass filter: 660 13 nm). In summary, we have investigated a FRET system with a multiexponentially decaying donor, multiple j141 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.6 FRET beyond F€orster j 5 How to Apply FRET: From Experimental Design to Data Analysis (a) (c) (b) (d) 0.1–0.9 ms Figure 5.17 Time-resolved Tb donor quenching (a and b) and QD acceptor sensitization (c and d). (a) Luminescence decay curves detected within the Tb donor channel for increasing QD acceptor concentrations: 0 nM (gray), 0.06, 0.1, and 0.15 nM (black from top to bottom). The white lines within the curves are the fitted curves. (b) Time-gated (0.1–0.9 ms) luminescence intensities detected within the Tb donor channel for increasing QD concentrations. (c) Luminescence decay 0.1–0.9 ms curves detected within the QD acceptor channel for increasing QD acceptor concentrations: 0 nM (gray), 0.06, 0.1, and 0.15 nM (black from bottom to top). The white lines within the curves are the fitted curves. (d) Time-gated (0.1–0.9 ms) luminescence intensities detected within the QD acceptor channel for increasing QD concentrations. (Adapted with permission from Ref. [56]. Copyright 2013, American Chemical Society.) D–A distances, and dynamic (distances) as well as static (free Tb concentration) quenching contributions using time-resolved detection of donor quenching and acceptor sensitization. The increasing QD acceptor concentration leads to Tb D-quenching and QD Asensitization, as shown in Figure 5.17 in the intensity decay curves as well as in the time-gated intensities. No background correction was performed for these graphs. The Tb emission cross talk to the QD acceptor detection channel can be seen in the lowest decay curve in Figure 5.17c and the time-gated intensity offset in Figure 5.17d. Due to the relatively low luminescence quantum yield of the QDs, the sensitized QD emission is relatively weak, but distinguishes significantly from the Tb background signal. The appearance of new short decay times due to FRET becomes clearly visible within both the Tb donor and the QD acceptor decay curves. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 142 The FRET-quenched donor curves can be conveniently fitted with triple-exponential decay functions (pure D was fitted with a double-exponential), whereas the FRET-sensitized acceptor curves require a quadruple-exponential fit function. Due to strong background fluorescence in the very short time range (strong saturated signals in the first 10–50 ms mainly due to sample autofluorescence and direct QD excitation) and the very low signal intensities for the acceptor channel in the very long time range (weak signal due to the Tb cross talk within the QD acceptor channel), the time ranges for the fits were chosen differently for D (0.02–8 ms) and A (0.05–4 ms). In most software tools for least-square fitting of exponential luminescence decays, such fits are called “tail fits,” which means that the fit starts at a different time t0 for D (t0 ¼ 0.02 ms) and A (t0 ¼ 0.05 ms). However, the complete luminescence decays of A and D start immediately after the excitation at 0 ms (the t0 values for the fits were chosen differently to improve the fit quality). Although such tail fits do not change the different single decay times, the single amplitudes (Ai-FIT) must be corrected (from the t0 fit values to the correct start of the luminescence decay at 0 ms) to yield the correct amplitudes (Ai) of the complete decay function: tt t t 0 0 ¼ Ai exp ) Ai ¼ AiFIT exp : I ¼ AiFIT exp t t t ð5:40Þ This is especially important when the amplitudes are necessary for the interpretation of the results (e.g., calculation of amplitude-averaged decay times or molecular fractions). For all fits of FRET-quenched and FRET-sensitized decay curves, the Tb donor decay time was a fixed value. The fitted curves are presented in Figure 5.17 and all fit results are presented in Table 5.5 for the Tb donor detection channel and in Table 5.6 for the QD acceptor detection channel. The triple-exponential FRET-quenched Tb donor decay curves were fitted for the amplitude fractions aDA 1, aDA 2, and aDA 0 and the decay times tDA1, tDA2, and tDA0, for which the third decay time component was fixed to tDA0 ¼ tD2 [the pure Tb donor has two decay times of tD1 ¼ (0.56 0.06) ms and tD2 ¼ (2.56 0.5) ms leading to an amplitude-averaged decay time of ktDi ¼ (2.27 0.5) ms] (cf. Tables 5.5 and 5.6) in order to take into account the emission of unquenched donors. For the calculation of the average donor decay time in the presence of the acceptor htDA i, only the first two amplitudes and decay times were used (as the third component represents unquenched donors). Therefore, the amplitude fractions must be redefined for these two decay times tDA1 and tDA2: aDA1 ¼ aDA 1 aDA 1 þ aDA 2 and aDA2 ¼ aDA 2 : aDA 1 þ aDA 2 ð5:41Þ As the unquenched donor possesses two decay time components (tD1 and tD2), htDA i must be corrected for the shorter time component (tD1). As the value of tD1 falls within the time range of the FRET-quenched decay times, the use of an additional exponential (with fixed tD1) for the fit procedure leads to inconsistent fit results. Therefore, a correction factor zD (the fraction of unquenched donors in the short time components) can be applied. zD is determined by comparing the j143 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.6 FRET beyond F€orster Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j 5 How to Apply FRET: From Experimental Design to Data Analysis amplitude fractions of tD2 and tDA0 (tDA0 ¼ tD2) multiplied by the amplitude fraction aD1: zD ¼ aD1 ðaDA 0 =aD2 Þ: ð5:42Þ The average FRET-quenched decay time is then (with aDA1 þ aDA2 ¼ 1) htDA i ¼ aDA1 tDA1 þ aDA2 tDA2 zD ðaDA1 þ aDA2 ÞtD1 aDA1 tDA1 þ aDA2 tDA2 zD tD1 ¼ aDA1 þ aDA2 zD ðaDA1 þ aDA2 Þ 1 zD ð5:43Þ and the average FRET efficiency hgFRET i is calculated by Equation 5.13 using the average decay times htDA i and htD i. The quadruple-exponential FRET-sensitized QD acceptor decay curves were fitted for the amplitude fractions aAD 1, aAD 2, aAD 3, and aAD 0 and the decay times tAD1, tAD2, tAD3, and tAD0, for which the fourth decay time component was fixed to tAD0 ¼ tD2 in order to take into account the emission of unquenched donors, which is much less intense compared to the donor channel, but still present due to spectral cross talk of the Tb emission in the QD acceptor detection channel. The correction factor zA (the fraction of unquenched donors in the short time components) is almost negligible, but is still taken into account for a correct treatment: zA ¼ aD1 ðaAD 0 =aD2 Þ: ð5:44Þ In order to calculate the average FRET decay time htAD i, only the amplitudes and lifetimes with i ¼ 1–3 are taken into account (i ¼ 0 represents the unquenched donor emission). Moreover, the amplitudes aAD i must be corrected by the FRET rates 1 (combination of Equations 5.8b and 5.13) to take into account kFRETi ¼ t1 ADi htD i the FRET efficiency-dependent excitation of the acceptors. The corrected amplitude fractions are (for i ¼ 1–3) aADi ¼ ðaAD ðaAD i =kFRETi Þ : 1 =kFRET1 Þ þ ðaAD 2 =kFRET2 Þ þ ðaAD 3 =kFRET3 Þ ð5:45Þ The average FRET decay time is then calculated by htAD i ¼ aAD1 tAD1 þ aAD2 tAD2 þ aAD3 tAD3 zA tD1 1 zA ð5:46Þ and the average FRET efficiency hgFRET i is calculated by Equation 5.13 using the average decay times htAD i (instead of tDA) and htD i. For each FRET decay time (from the donor and the acceptor fits), a specific D–A distance r can be calculated using Equation 5.14. The fractions of FRET pairs found at the different distances corresponding to tDAi and tADi are given by the amplitude fractions of these decay times. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 146 5.6.2 BRET and CRET The discovery of bioluminescence (of bacteria) [57–59] and chemiluminescence (of organic compounds) [60,61] dates back to the nineteenth century. For both phenomena, luminescence is created by a chemical reaction (as excitation source) inside (bioluminescence) or outside (chemiluminescence) a living organism. If the chemically excited molecules are used as energy donors in combination with a suitable acceptor inside a FRET system, the energy transfer is called bioluminescence resonance energy transfer and chemiluminescence resonance energy transfer, respectively. As only the donor excitation is different from FRET and the energy transfer mechanism is the same, the FRET theory (r6 distance dependence, etc.) can be applied for BRET and CRET. The first investigations of BRET and CRET date back to approximately half a century ago [62–64], but BRET and CRET applications have experienced a recent renaissance, especially due to novel FRET acceptors such as fluorescent proteins and nanoparticles [65–67]. Most BRET applications use luciferases (e.g., Renilla luciferase “Rluc” or Firefly luciferase “Fluc”), which catalyze the oxidation of their substrates (e.g., coelenterazine for Rluc or luciferin for Fluc), to produce emission in the blue to green spectral range (emission peaks of about 480 nm for coelenterazine and 570 nm for luciferin) for acting as BRET donors. In CRET, mainly luminol derivatives are used as donors. Luminol oxidation is a nonenzymatic reaction with much lower efficiency than the bioluminescence reactions [68]. Both BRETand CRETcan be used with different acceptors such as organic dyes, fluorescent proteins, or quantum dots. Figure 5.18 shows two recent examples of BRETand CRET. The BRETsystem presents hybrid molecules consisting of a firefly luciferin donor and different organic dye acceptors [69]. This approach allowed the development of luciferins emitting in the near-infrared, which is an important wavelength range because of deeper tissue penetration compared to UV or visible light. Moreover, these novel NIR luciferins did not require any ex vivo luciferase manipulation. Although the brightness of the BRET luminescence was very low, the authors could show NIR luminescence in live cells and living mice using luciferase-expressing cells. The CRET system is embedded in a DNA machine [70], for which a nucleic acid scaffold (1) is hybridized with three DNA footholds. The first foothold is labeled with a FAM dye (2) or a semiconductor quantum dot (3). The second foothold (4) is initially free and the third foothold (5) is hybridized to a nucleic acid strand (6) that acts as DNA walker and contains a hemin/G-quadruplex DNAzyme sequence (caged in the duplex structure with the third foothold). Addition of a fuel strand (7) leads to strand displacement of the walker (6) by the formation of a more stable 7/5 duplex and the hybridization of 6 overhang to the second foothold (4). In this configuration, 6 can form a hemin/Gquadruplex DNAzyme, which catalyzes the generation of chemiluminescence in the presence of luminol and H2O2. The activated luminol is acting as a CRETdonor for the FAM dye (2) or quantum dot (3), which then produces its own luminescence. Notably, this process is completely reversible by the addition of an antifuel strand (8), leading to a strand replacement of 7 to form a stable 7/8 duplex and the reverse walking step of 6 to 5, which switches the chemiluminescence (and the CRET) back to “off.” j147 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.6 FRET beyond F€orster j 5 How to Apply FRET: From Experimental Design to Data Analysis Figure 5.18 Recent examples of BRET (top) and CRET (bottom) systems. (a) Emission spectra of the pure donor aminoluciferin (AL) and the acceptor–donor complexes Cy5-AL, SiR700-AL, and Cy7-AL, which give access to the NIR wavelength range. (b) Detection of BRET from AL to Cy7 in luciferase-expressing cells injected to mice. In contrast to the BRET emission of Cy7-AL (bottom), the pure AL sample (top) does not show any Cy7 luminescence. (c) Creation of CRET from activated luminol to a FAM dye or a quantum dot (on 2 and 3, respectively) by switchable hemin/G-quadruplex formation. The DNA machine is switched “ON” by 7, which leads to a walkover of 6 to 4 and the generation of CRET followed by light emission of FAM or QD. Addition of 8 switches the machine back to “OFF” because of a stable formation of a 7/8 duplex and the reverse walk of 6 to 5, causing the extinction of CRET. (d) Luminol chemiluminescence (large peak around 420 nm) and CRET-sensitized FAM luminescence (small peak around 518 nm). The inset shows the switchable CRET signals of the DNA walker system. (e) Luminol chemiluminescence (peak around 420 nm) and CRET-sensitized QD luminescence (peak around 615 nm). The inset shows the switchable CRET signals of the DNA walker system. (Parts (a) and (b) reprinted with permission from Ref. [69]. Copyright 2013, Wiley-VCH Verlag GmbH. Parts (c–e) reprinted with permission from Ref. [70]. Copyright 2012, American Chemical Society.) 5.6.3 Energy Transfer to Metal Nanoparticles (FRET, NSET, DMPET, NPILM, etc.) The good news about distance-dependent energy transfer to metal (mainly Au and in some cases Ag is used) nanoparticles or nanoclusters is that it works very efficiently. However, different mechanisms have been proposed to be responsible for the Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 148 distance dependence, which has been investigated in theory [71–76] and in practice for different energy donors such as organic dyes [77–84], semiconductor quantum dots [85–90], or fluorescent proteins [91]. Among all these different systems, different mechanisms have been proposed to be responsible for the energy transfer and a general model for energy transfer to metal nanoparticles does not exist. The two most important aspects for influencing the distance dependence are the donor– acceptor distance and the nanoparticle size. At small distances (separation between donor and acceptor is smaller than the size of the donor and/or the acceptor), nonradiative energy transfer is proposed to be the main cause of quenching; whereas at larger distances, radiative energy transfer will play a major role. One of the main ideas for short distances is that the point dipole approximation is not valid anymore leading to a deviation in the FRET distance dependence by overestimating the FRET rate. One can start from a generalization of Equation 5.7 for resonance energy transfer (RET): n D0 kRET ¼ t1 ; ð5:47Þ D d where D0 is the donor–acceptor distance for 50% energy transfer efficiency (R0 in FRET) and d is the donor–acceptor separation distance (r in FRET). The main difference between the RET theories can be found in the exponent n, which is n ¼ 6 for FRET. Agreement of the FRET theory with experimental data was shown for Au nanoparticles with diameters of 1.4 nm [85], 5 nm [89], and 15 and 80 nm [87], all using quantum dots as donors. The DMPET model of Carminati et al. [72] takes into account the distance dependence of radiative (mainly n ¼ 3 with a n ¼ 6 contribution at plasmon resonance) and nonradiative (n ¼ 6) decay rates, which leads to some additional correction terms compared to FRET [88]. Moreover, nonradiative decay is strongly enhanced, when the donor radiates at the plasmon resonance wavelength of the nanoparticle [72]. Moroz reanalyzed Carminati’s DMPET model of a 10 nm diameter Ag nanoparticle and pointed out that there is a significant contribution (between 50 and 101% of the total value) of higher order multipoles to nonradiative rates even at 5 nm donor–acceptor distance [75]. This theoretical model provided very good agreement with luminescence decay time quenching of quantum dots by Au nanoparticles of 10, 15, and 20 nm diameter positioned at distances of about 17, 15, and 13 nm from the quantum dots using DNA origami [86]. The NSET model proposed by Strouse and coworkers [79,80,82–84] uses a d4 distance dependence (n ¼ 4), which significantly increases (about twofold) the distance range of FRET. They proposed the following NSET transfer rate: kNSET ¼ 0:225 c3 v2D vF kF d4 WD ; tD ð5:48Þ where c is the speed of light, WD is the donor quantum yield, vD is the angular frequency (v ¼ 2pcl1) for the donor, vF is the angular frequency for bulk gold, and kF is the Fermi vector for bulk gold. NSET has mainly been found to be in good agreement with experimental data (using organic dyes and quantum dots as donors) if the j149 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.6 FRET beyond F€orster j 5 How to Apply FRET: From Experimental Design to Data Analysis Figure 5.19 (a) Distance dependence of the quenching efficiency for different sizes of Au nanoparticles. The 8 nm Au nanoparticle was in good agreement with NSET theory, whereas the other two were attributed to radiative energy transfer and could not be fitted with FRET, NSET, or DMPET. (b) Increase of the 50% energy transfer efficiency value (R0) with the size of gold nanoparticle. (Reprinted with permission from Ref. [78]. Copyright 2009, Wiley-VCH Verlag GmbH.) nanoparticles are of small size (below about 3 nm diameter) and thus do not have any plasmon bands [79,80,82,84,87,88]. Nevertheless, also larger Au nanoparticles of up to 18 nm diameters showed experimental NSET behavior [77,78,90,91]. Bhowmick et al. proposed a theoretical model with a n ¼ 6 distance dependence of energy transfer to surface plasmonic modes at large separation between a dye and a nanoparticle and a 3 < n < 4 distance dependence of energy transfer for short separation (similar to the nanoparticle size) between dye and nanoparticle [71]. As already mentioned, for large distances the energy transfer is mainly governed by the radiative rate with a d3 distance dependence (n ¼ 3), leading to a large energy transfer distance range compared to FRET. Indeed, such energy transfer between dyes and Au nanoparticles over distances of more than 40 nm was found by steadystate experiments [78] and decay time measurements [81]. In the latter publication, the acronym NPILM was introduced. Figure 5.19 shows experimental data of energy transfer over distances of up to 50 nm for different nanoparticle sizes. 5.6.4 Other Transfer Mechanisms Apart from the FRET-like energy transfer mechanisms mentioned so far, there are several other energy or charge transfer mechanisms that can enlarge the distance range of FRET in both the short and the long directions. On the short end, there are electron exchange (Dexter) or electron transfer (Marcus) mechanisms related to orbital overlap, with exponentially decaying distance dependence. On the long end, there are mechanisms such as plasmon coupling (up to about 80 nm donor–acceptor distance) or singlet oxygen diffusion (up to about 200 nm). Although all of these mechanisms are not based on nonradiative Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 150 Table 5.7 Overview of different distance-dependent energy/charge transfer mechanisms. Transfer mechanism Dexter transfer Charge transfer FRET/BRET/ CRET/DMPET NSET Plasmon coupling Singlet oxygen transfer Distance rangea) Below about 1 nm Distance dependence Comments exp(-r) (Equation 5.49) exp(-r) (Equation 5.50) 1/r6 (Equation 5.7) Energy transfer by electron exchange Below about 2 nm Electron or hole transfer (Marcus theory) Energy transfer without About 1–20 nm electron exchange About 1–40 nm 1/r4 (Equation 5.48) FRET from a donor to a metal surface About 5–80 nm (up to exp(-r) Size, shape, material, and 300 nm in theory) (Equation 5.52) medium-dependent wavelength shift About 10–100 nm (up to exp(-r) Not used for distance 250 nm in theory) (Equation 5.53) measurements a) All values refer to single-step transfer (one donor–acceptor pair). Energy migration or electron/hole hopping can lead to larger overall transfer distances. energy transfer due to dipole–dipole interactions and are thus not directly related to FRET, they will be briefly described here (the interested reader is referred to further literature within the following sections) in order to give a broader picture of distance-dependent energy/charge transfer mechanisms. The different mechanisms, their distance dependence, and their approximate distance range are shown in Table 5.7. 5.6.4.1 Electron Exchange Energy Transfer (Dexter Transfer) In the case of overlapping orbitals of donor and acceptor molecules, which require short D–A distances, electron exchange between D and A can occur. This mechanism is different from the Coulombic interaction in FRET or the electron tunneling in charge transfer (Figure 5.20). The electron exchange rate is related to the orbital overlap, which is expected to fall off exponentially with increasing D–A distance. Electron exchange requires energetic resonance of D and A and therefore the exchange rate will also be dependent on the spectral overlap of D and A. A theory for electron exchange-mediated energy transfer was developed by Dexter in 1953 [92]. The rate constant of electron exchange energy transfer [or Dexter transfer (DT)] is given by 2r kDT ¼ KJ DT exp ; ð5:49Þ L where K is a constant related to specific orbital interactions and JDT is the spectral overlap integral. This spectral overlap is similar to the J in FRET (cf. Equation 5.9) with the important difference that both the fluorescence and the absorption j151 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.6 FRET beyond F€orster j 5 How to Apply FRET: From Experimental Design to Data Analysis Figure 5.20 Different mechanisms for generating a ground-state donor and an excited acceptor (D þ A as shown in the center) by FRET (top left: Coulombic coupling of D and A), Dexter transfer (DT, bottom left: electron exchange between D and A and A and D) or charge transfer (top right, CT: D as electron donor; bottom right, CTþ: D as hole donor). spectrum are normalized. This means that (in contrast to FRET) the spectral overlap integral is not dependent on the molar absorptivity (extinction coefficient). r is the edge-to-edge separation between D and A and L is the sum of their van der Waals radii. As the transfer rate of Dexter transfer decreases exponentially with r, kDT becomes negligibly small for D–A distances of more than one or two molecular diameters (about 0.5–1 nm). In contrast to charge transfer (next section), the solvent plays a minor role (apart from establishing the collision of D and A via diffusion) for the transfer. As the constant K cannot be easily related to experimentally determinable quantities, it is difficult to perform a quantitative experimental characterization of Dexter transfer. More details about the Dexter theory can be found in photochemistry and spectroscopy textbooks [17,18]. 5.6.4.2 Charge Transfer (Marcus Theory) Similar to Dexter transfer, charge transfer requires orbital overlap and has therefore exponential distance dependence. The main difference from the Dexter electron exchange lies in the transfer mechanism, as illustrated in Figure 5.20. Electron exchange is a concerted two-electron transfer, whereas electron transfer (or charge separation) requires an electron (or hole) donor and acceptor. Although charge transfer can occur between molecules in the energetic ground state, an excited molecule is both a better reducing agent (or reductant) and a better oxidizing agent (or oxidant) due to a lower ionization potential and a larger electron affinity, respectively [17]. The tunneling of electrons is dependent on the reactants and the solvent requiring a reorganization of both, which can be divided into inner sphere and outer sphere reorganization. The theory of this distance-dependent charge transfer was developed by Marcus in 1956 [93–95]. The rate of charge transfer (CT) can be written as Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 152 kCT ¼ 2p 2 exp ½bðr r 0 Þ ðDGðrÞ þ lðrÞÞ2 ; J 0 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp 4lðrÞkB T h 4plðrÞkB T ð5:50Þ where r is the center-to-center distance between D and A, r0 is the distance for which D and A are in contact, b characterizes the distance dependence of the coupling, J0 is the contact value of the donor–acceptor electronic coupling matrix element, l(r) is the distance-dependent reorganization energy, DG(r) is the distance-dependent free energy change, and h and kB are the reduced Planck’s constant and the Boltzmann constant, respectively [96]. Equation 5.50 contains many distance-dependent parameters, which shows that a correct treatment of distance dependence can be complicated and simple exponential models have been applied to fit experimental data (e.g., for the distance dependence of electron transfer in DNA): kCT ¼ AET exp ðbET ðr r 0 ÞÞ; ð5:51Þ where AET is a preexponential factor and bET characterizes the distance dependence of the transfer. This short section can only give a first glance into the charge transfer mechanism. The most important aspects in relation to FRET (the main topic of this book) are the short distance range (which can take higher values than Dexter transfer due to the different mechanism of charge transfer) of up to about 2 nm or even higher values (in the case of electron hopping, over several redox centers) [97,98] and the exponential distance dependence. Electron (or charge) transfer has been intensively studied for chemical and biological systems and details can be found in textbooks and review articles [17,99–115]. Although the D–A distance is limited, charge transfer has the advantage that it does not require any spectral overlap (in contrary to FRET or Dexter transfer) and therefore one suitable electron donor (or acceptor) can be used to quench several different luminescent molecules for multiplexing purposes. Different biological and chemical sensing concepts using luminescence quenching of quantum dots by different charge transfer agents have been recently developed [116–123]. Such charge transfer sensors allowed the analysis of simultaneous quenching of eight different semiconductor quantum dots [124]. 5.6.4.3 Plasmon Coupling Upon interaction with light, noble metal nanoparticles (most often Au and Ag are used) can display localized surface plasmon resonance (LSPR) leading to broad and strong absorption or scattering bands in the UV-Vis wavelength region. These unique optical properties can be used for many different biosensing applications [125,126]. The resonant frequency (or wavelength) of metal nanoparticles depends on their material, size, shape, and surrounding medium. When two such LSPR nanoparticles are brought into proximity, their plasmon resonances can couple, which results in redshifted absorption or scattering bands of the two coupled particles. This wavelength shift is dependent on the particle separation and can therefore be used as spectroscopic or plasmonic ruler [127–133]. The distancedependent wavelength shift decays approximately exponentially and covers a large j153 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.6 FRET beyond F€orster j 5 How to Apply FRET: From Experimental Design to Data Analysis distance range. Interparticle distances of up to about 75 nm have been measured [130] and theoretical calculations predicted wavelength shifts for separations of up to about 300 nm [132]. Jain et al. derived an empirical plasmon ruler equation for the distance dependence of the wavelength shift, which is applicable for different particle sizes [128]: Dl A exp l0 d=D ; t ð5:52Þ where Dl/l0 is the fractional wavelength shift, A is the magnitude of the fractional shift, d is the interparticle edge-to-edge separation, D is the particle diameter, and t is the decay constant of the exponential decay. The authors found that t is similar for different particle materials, sizes, shapes, and medium electric constant (these parameters change the amplitude A) and takes a value close to t ¼ 0.23. Although the distance range of plasmon rulers is significantly larger than for FRET and there is no dependence on the relative orientation of D and A, there are several limitations of this technique for absolute distance measurements mainly due to size and shape inheterogeneity of the nanoparticles, the relatively large size of the nanoparticles in order to achieve a strong and sensitive scattering signal, and the dependence of plasmon resonance on the refractive index of the medium [130,131]. This means (although an empirical plasmon ruler equation exists) that each plasmon ruler must be carefully calibrated before its application for the determination of distances in unknown systems. Moreover, plasmon rulers cannot provide the inherent ratiometric behavior of FRET (D-quenching and A-sensitization) and are limited in multiplexed detection. 5.6.4.4 Singlet Oxygen Diffusion Another possibility to transfer energy over larger distances is to use singlet oxygen diffusion. This technique was developed by Ullman et al. and commercialized under the brand name LOCI1 (Luminescent Oxygen Channeling Immunoassay, Behring Diagnostics Inc.) [134,135]. The energy transfer is based on the following principle: A nanoparticle charged with a photosensitizer (phthalocyanine) produces singlet oxygen upon light excitation around 680 nm. The singlet oxygen can diffuse to a nearby second nanoparticle that is charged with dioxene (or thioxene) that produces chemiluminescence upon reaction with singlet oxygen. As the quantum yield of this chemiluminescence is very low, an additional fluorophore [9,10-bisphenylethynylanthracene (BPEA) or Eu(TTA)3Phen] is added, which results in a much higher overall quantum yield. A homogeneous assay format, for example, two LOCI-nanoparticlelabeled antibodies binding to a specific biomarker, can fix both nanoparticles at a distance, for which singlet oxygen can diffuse from particle one to particle two. The produced chemiluminescence intensity is then proportional to the biomarker concentration. The same principle is used today in the commercial biosensing platform AlphaLISA1 by Perkin Elmer [136] and the detection of different biomarkers using this technology can be found in the literature [137–141]. As the concentration of singlet oxygen, which can generate chemiluminescence within the second nanoparticle, is Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 154 dependent on diffusion, the distance dependence of the energy transfer (or the single oxygen concentration) can be calculated by Fick’s first law and is given by [134] cSO # ra pffiffiffiffiffiffiffiffiffiffiffi exp pffiffiffiffiffiffiffiffiffiffiffi ; ¼ 4prDð1 þ a= D=kD Þ D=kD s " ð5:53Þ where cSO is the concentration of singlet oxygen at a distance r from the center of the first nanoparticle, S is the rate of singlet oxygen formation by the particle, a is the particle radius, kD is the singlet oxygen decay constant in water, and D is the diffusion coefficient in water. Due to the relatively large nanoparticles (150 nm diameter), which need to contain high amounts of sensitizers and chemiluminescence compounds for the generation of intense luminescence signals, this technology is not suited for distance measurements (or at least – to my knowledge – it has not yet been tried out). Similar to plasmon coupling, LOCI is not inherently ratiometric and needs to be carefully calibrated. 5.7 Summary and Outlook In summary, FRET is a very powerful technique for the measurement of distances and concentrations with very high precision and sensitivity on a length scale of about 1–20 nm. F€orster’s theory for the relation between spectroscopic data and the FRET distance dependence dates back to 1946 and therefore FRET is probably one of the first optical superresolution techniques. Thanks to the development of many types of fuorophores over the last decades, there are numerous possibilities of choosing an adequate donor–acceptor pair (cf. Chapters 6 and 14) for the nanometric system of interest (cf. Chapters 6–13). Before planning a FRET experiment, one should also carefully think about the expected distances and the (biological) recognition mechanism in which donor and acceptor will be involved. FRET can be characterized by different technologies using luminescence quantum yields, intensities, and lifetimes and both the donor (quenching or photobleaching) and the acceptor (sensitization or photobleaching) can be analyzed in order to achieve accurate results. When planning the experiments as well as when analyzing the results, one should always have in mind that there are other energy or charge transfer mechanisms that can be responsible for the quenching of a donor and/or the sensitization of an acceptor, and that there are complementary techniques to enlarge the distance range of FRET. The design of novel energy transfer concepts and probes by efficient combination using the large variety of fluorophores and technologies will make FRET (and the other energy/charge transfer technologies) an equally or even more important and powerful technology in the future. FRET on, FRET jocks! j155 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 5.7 Summary and Outlook j 5 How to Apply FRET: From Experimental Design to Data Analysis Acknowledgment I would like to thank Dr. Daniel Geißler for his comments on and review of this chapter. References 1 H€ otzer, B., Medintz, I.L., and 2 3 4 5 6 7 8 9 10 11 Hildebrandt, N. (2012) Fluorescence in nanobiotechnology: sophisticated fluorophores for novel applications. Small, 8 (15), 2297–2326. Clegg, R.M. (1996) Fluorescence resonance energy transfer, in Fluorescence Imaging Spectroscopy and Microscopy (eds X.F. Wang and B. Herman), John Wiley & Sons, Inc., New York, pp. 179–251. Clegg, R.M. (2009) F€orster resonance energy transfer: FRET what is it, why do it, and how it’s done, in Laboratory Techniques in Biochemistry and Molecular Biology (ed. T.W.J. Gadella), Academic Press, Burlington, pp. 1–57. Gadella, T.W.J. (2009) FRET and FLIM Techniques, Elsevier, Amsterdam. Jares-Erijman, E.A. and Jovin, T.M. (2003) FRET imaging. Nature Biotechnology, 21 (11), 1387–1395. Jares-Erijman, E.A. and Jovin, T.M. (2006) Imaging molecular interactions in living cells by FRET microscopy. Current Opinion in Chemical Biology, 10 (5), 409–416. Morrison, L.E. (1988) Time-resolved detection of energy-transfer: theory and application to immunoassays. Analytical Biochemistry, 174 (1), 101–120. Sekar, R.B. and Periasamy, A. (2003) Fluorescence resonance energy transfer (FRET) microscopy imaging of live cell protein localizations. Journal of Cell Biology, 160 (5), 629–633. Selvin, P.R. (2000) The renaissance of fluorescence resonance energy transfer. Nature Structural Biology, 7 (9), 730–734. Stryer, L. (1978) Fluorescence energy transfer as a spectroscopic ruler. Annual Review of Biochemistry, 47, 819–846. Van der Meer, B.W., Coker, G., and Chen, S.Y.S. (1994) Resonance Energy Transfer: 12 13 14 15 16 17 18 19 20 21 Theory and Data, John Wiley & Sons, Inc., New York. Vogel, S.S., Thaler, C., and Koushik, S.V. (2006) Fanciful FRET. Science’s STKE: Signal Transduction Knowledge Environment, 2006 (331), re2. Wu, P.G. and Brand, L. (1994) Resonance energy-transfer methods and applications. Analytical Biochemistry, 218 (1), 1–13. Sapsford, K.E., Berti, L., and Medintz, I.L. (2006) Materials for fluorescence resonance energy transfer analysis: beyond traditional donor–acceptor combinations. Angewandte Chemie, International Edition, 45 (28), 4562–4588. Lakowicz, J.R. (2006) Principles of Fluorescence Spectroscopy;, 3rd edn, Springer, New York, NY. Periasamy, A. and Day, R.N. (2005) Molecular Imaging: FRET Microscopy and Spectroscopy;, Oxford University Press, New York. Turro, N.J., Ramamurthy, V., and Scaiano, J.C. (2010) Modern Molecular Photochemistry of Organic Molecules;, University Science Books, Sausalito, Calif. Valeur, B. (2002) Molecular Fluorescence: Principles and Applications;, Wiley-VCH Verlag GmbH, Weinheim. Andrews, D.L. and Demidov, A.A. (1999) Resonance Energy Transfer, John Wiley & Sons, Ltd, Chichester. Braslavsky, S.E., Fron, E., Rodriguez, H.B., Roman, E.S., Scholes, G.D., Schweitzer, G., Valeur, B., and Wirz, J. (2008) Pitfalls and limitations in the practical use of Forster’s theory of resonance energy transfer. Photochemical & Photobiological Sciences, 7 (12), 1444–1448. Clegg, R.M. (2006) The history of FRET: from conception through the labors of birth, in Reviews in Fluorescence 2006, vol. 3 (eds C.D. Geddes and J.R. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 156 22 23 24 25 26 27 28 29 30 31 32 Lakowicz), Springer, New York, p. 1 (online resource). F€ orster, T. (1946) Energiewanderung und fluoreszenz. Naturwissenschaften, 33 (6), 166–175. F€ orster, T. (1948) Zwischenmolekulare energiewanderung und fluoreszenz. Annalen der Physik, 2 (1–2), 55–75. F€ orster, T. (1949) Experimentelle und theoretische untersuchung des € bergangs von zwischenmolekularen u elektronenanregungsenergie. Zeitschrift f€ ur Naturforschung Section A: Journal of Physical Sciences, 4 (5), 321–327. F€ orster, T. (1951) Fluoreszenz organischer Verbindungen;, Vandenhoeck & Ruprecht, Gottingen. F€ orster, T. (1959) 10th Spiers Memorial Lecture: transfer mechanisms of electronic excitation. Discussions of the Faraday Society, 27, 7–17. F€ orster, T. (1965) Delocalized excitation and excitation transfer, in Modern Quantum Chemistry, Istanbul Lectures: Part III: Action of Light and Organic Crystals (ed. O. Sinanoglu), Academic Press, New York, pp. 93–137. Clegg, R.M. (1992) Fluorescence resonance energy-transfer and nucleicacids. Methods in Enzymology, 211, 353–388. Clegg, R.M., Murchie, A.I.H., Zechel, A., Carlberg, C., Diekmann, S., and Lilley, D.M.J. (1992) Fluorescence resonance energy-transfer analysis of the structure of the 4-way DNA junction. Biochemistry, 31 (20), 4846–4856. Clegg, R.M., Murchie, A.I.H., Zechel, A., and Lilley, D.M.J. (1993) Observing the helical geometry of double-stranded DNA in solution by fluorescence resonance energy-transfer. Proceedings of the National Academy of Sciences of the United States of America, 90 (7), 2994–2998. Jovin, T.M. and Arndtjovin, D.J. (1989) Luminescence digital imaging microscopy. Annual Review of Biophysics and Biophysical Chemistry, 18, 271–308. Raicu, V. (2007) Efficiency of resonance energy transfer in homo-oligomeric complexes of proteins. Journal of Biological Physics, 33 (2), 109–127. 33 Clapp, A.R., Medintz, I.L., Mauro, J.M., 34 35 36 37 38 39 40 41 42 43 Fisher, B.R., Bawendi, M.G., and Mattoussi, H. (2004) Fluorescence resonance energy transfer between quantum dot donors and dye-labeled protein acceptors. Journal of the American Chemical Society, 126 (1), 301–310. Berney, C. and Danuser, G. (2003) FRET or no FRET: a quantitative comparison. Biophysical Journal, 84 (6), 3992–4010. Corry, B., Jayatilaka, D., and Rigby, P. (2005) A flexible approach to the calculation of resonance energy transfer efficiency between multiple donors and acceptors in complex geometries. Biophysical Journal, 89 (6), 3822–3836. Kenworthy, A.K. (2005) Photobleaching FRET microscopy, in Molecular Imaging: FRET Microscopy and Spectroscopy (eds A. Periasamy and R.N. Day), Oxford University Press, New York, pp. 146–164. Charbonniere, L.J. and Hildebrandt, N. (2008) Lanthanide complexes and quantum dots: a bright wedding for resonance energy transfer. European Journal of Inorganic Chemistry, 2008 (21), 3241–3251. Sillen, A. and Engelborghs, Y. (1998) The correct use of “average” fluorescence parameters. Photochemistry and Photobiology, 67 (5), 475–486. Selvin, P.R. (1995) Fluorescence resonance energy-transfer. Methods in Enzymology, 246, 300–334. Selvin, P.R. (1996) Lanthanide-based resonance energy transfer. IEEE Journal of Selected Topics in Quantum Electronics, 2 (4), 1077–1087. Selvin, P.R. (2002) Principles and biophysical applications of lanthanidebased probes. Annual Review of Biophysics and Biomolecular Structure, 31, 275–302. Selvin, P.R. and Hearst, J.E. (1994) Luminescence energy transfer using a terbium chelate: improvements on fluorescence energy transfer. Proceedings of the National Academy of Sciences of the United States of America, 91 (21), 10024–10028. Selvin, P.R., Rana, T., Klein, M.P., and Hearst, J.E. (1993) Dark-background fluorescence energy-transfer using j157 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 5 How to Apply FRET: From Experimental Design to Data Analysis 44 45 46 47 48 49 50 51 52 53 lanthanide chelates. Biophysical Journal, 64 (2), A243–A243. Selvin, P.R., Rana, T.M., and Hearst, J.E. (1994) Luminescence resonance energytransfer. Journal of the American Chemical Society, 116 (13), 6029–6030. Xiao, M. and Selvin, P.R. (2001) Quantum yields of luminescent lanthanide chelates and far-red dyes measured by resonance energy transfer. Journal of the American Chemical Society, 123 (29), 7067–7073. B€ unzli, J.-C.G. (1989) Luminescent probes, in Lanthanide Probes in Life, Chemical, and Earth Sciences: Theory and Practice (eds J.-C.G. B€ unzli and G.R. Choppin), Elsevier, Amsterdam. B€ unzli, J-.C.G. (2004) Luminescent lanthanide probes as diagnostic and therapeutic tools, in Metal Ions in Biological Systems, Vol 42: Metal Complexes in Tumor Diagnosis and as Anticancer Agents, vol. 42 (eds A. Sigel and H. Sigel), Marcel Dekker Inc., New York, pp. 39–75. B€ unzli, J.C.G. (2010) Lanthanide luminescence for biomedical analyses and imaging. Chemical Reviews, 110 (5), 2729–2755. B€ unzli, J.C.G., Chauvin, A.S., Kim, H.K., Deiters, E., and Eliseeva, S.V. (2010) Lanthanide luminescence efficiency in eight- and nine-coordinate complexes: role of the radiative lifetime. Coordination Chemistry Reviews, 254 (21–22), 2623–2633. Eliseeva, S.V. and B€ unzli, J.C.G. (2010) Lanthanide luminescence for functional materials and biosciences. Chemical Society Reviews, 39 (1), 189–227. Geißler, D. and Hildebrandt, N. (2011) Lanthanide complexes in FRET applications. Current Inorganic Chemistry, 1, 17–35. Richardson, F.S. (1982) Terbium(III) and Europium(III) ions as luminescent probes and stains for biomolecular systems. Chemical Reviews, 82 (5), 541–552. Geißler, D., Charbonniere, L.J., Ziessel, R.F., Butlin, N.G., L€ohmannsr€oben, H.G., and Hildebrandt, N. (2010) Quantum dot biosensors for ultrasensitive multiplexed diagnostics. Angewandte Chemie, International Edition, 49 (8), 1396–1401. 54 Geißler, D., Stufler, S., L€ ohmannsr€oben, 55 56 57 58 59 60 61 62 63 H.G., and Hildebrandt, N. (2013) Six-color time-resolved F€orster resonance energy transfer for ultrasensitive multiplexed biosensing. Journal of the American Chemical Society, 135 (3), 1102–1109. Mathis, G. (1993) Rare-earth cryptates and homogeneous fluoroimmunoassays with human sera. Clinical Chemistry, 39 (9), 1953–1959. Wegner, K.D., Lanh, P.T., Jennings, T., Oh, E., Vaibhav, J., Fairclough, S.M., Smith, J.M., Giovanelli, E., Lequeux, N., Pons, T., and Hildebrandt, N. (2013) Influence of luminescence quantum yield, surface coating, and functionalization of quantum dots on the sensitivity of time-resolved FRET bioassays. ACS Applied Materials and Interfaces, 5 (8), 2881–2892. doi: dx.doi. org/10.1021/am3030728 Beyerinck, M.W. (1889) Le photobacterium luminosum, bacterie lumineuse de la mer du nord. Archives Neederlandaises des Sciences Exactes Naturelles, 13, 401–415. Giard, A. (1889) Sur l’infection phosphorescente des talitres et autres crustaces. Comptes rendus hebdomadaires des seances de l’Academie des Sciences, 7, 503–506. Nealson, K.H. and Hastings, J.W. (1979) Bacterial bioluminescence: its control and ecological significance. Microbiological Reviews, 43 (4), 496–518. Radziszewski, B. (1877) Untersuchungen €ber hydrobenzamid, amarin und lophin. u Berichte der Deutschen Chemischen Gesellschaft zu Berlin, 10, 70–75. Trautz, M. (1908) Chemilumineszenz. Zeitschrift f€ ur Elektrochemie und Angewandte Physikalische Chemie, 14 (33), 453–460. Johnson, F.H., Shimomura, O., Saiga, Y., Gershman, L.C., Reynolds, G.T., and Waters, J.R. (1962) Quantum efficiency of cypridina luminescence, with a note on that of aequorea. Journal of Cellular and Comparative Physiology, 60 (1), 85–103. Morin, J.G. and Hastings, J.W. (1971) Energy transfer in a bioluminescent system. Journal of Cellular Physiology, 77 (3), 313–318. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 158 64 White, E.H. and Roswell, D.F. (1967) 65 66 67 68 69 70 71 72 73 74 Intramolecular energy transfer in chemiluminescence. Journal of the American Chemical Society, 89 (15), 3944–3945. Huang, X.Y. and Ren, J.C. (2012) Nanomaterial-based chemiluminescence resonance energy transfer: a strategy to develop new analytical methods. Trends in Analytical Chemistry, 40, 77–89. Pfleger, K.D.G. and Eidne, K.A. (2006) Illuminating insights into protein–protein interactions using bioluminescence resonance energy transfer (BRET). Nature Methods, 3 (3), 165. Xia, Z.Y. and Rao, J.H. (2009) Biosensing and imaging based on bioluminescence resonance energy transfer. Current Opinion in Biotechnology, 20 (1), 37–44. Grayeski, M.L. (1987) Chemiluminescence analysis. Analytical Chemistry, 59 (21), A1243. Kojima, R., Takakura, H., Ozawa, T., Tada, Y., Nagano, T., and Urano, Y. (2013) Rational design and development of nearinfrared-emitting firefly luciferins available in vivo. Angewandte Chemie, International Edition, 52 (4), 1175–1179. Liu, X.Q., Niazov-Elkan, A., Wang, F.A., and Willner, I. (2013) Switching photonic and electrochemical functions of a DNAzyme by DNA machines. Nano Letters, 13 (1), 219–225. Bhowmick, S., Saini, S., Shenoy, V.B., and Bagchi, B. (2006) Resonance energy transfer from a fluorescent dye to a metal nanoparticle. Journal of Chemical Physics, 125 (18), 181102. Carminati, R., Greffet, J.J., Henkel, C., and Vigoureux, J.M. (2006) Radiative and non-radiative decay of a single molecule close to a metallic nanoparticle. Optics Communications, 261 (2), 368–375. Klimov, V.V., Ducloy, M., and Letokhov, V.S. (1996) Radiative frequency shift and linewidth of an atom dipole in the vicinity of a dielectric microsphere. Journal of Modern Optics, 43 (11), 2251–2267. Klimov, V.V., Ducloy, M., and Letokhov, V.S. (2001) Spontaneous emission of an atom in the presence of nanobodies. Quantum Electronics, 31 (7), 569–586. 75 Moroz, A. (2010) Non-radiative decay of a 76 77 78 79 80 81 82 83 dipole emitter close to a metallic nanoparticle: importance of higher-order multipole contributions. Optics Communications, 283 (10), 2277–2287. Persson, B.N.J. and Lang, N.D. (1982) Electron–hole-pair quenching of excited states near a metal. Physical Review B, 26 (10), 5409–5415. Chen, Y., O’Donoghue, M.B., Huang, Y.-F., Kang, H., Phillips, J.A., Chen, X., Estevez, M.C., Yang, C.J., and Tan, W. (2010) A surface energy transfer nanoruler for measuring binding site distances on live cell surfaces. Journal of the American Chemical Society, 132 (46), 16559–16570. Griffin, J., Singh, A.K., Senapati, D., Rhodes, P., Mitchell, K., Robinson, B., Yu, E., and Ray, P.C. (2009) Size- and distance-dependent nanoparticle surfaceenergy transfer (NSET) method for selective sensing of hepatitis C virus RNA. Chemistry: A European Journal, 15 (2), 342–351. Jennings, T.L., Schlatterer, J.C., Singh, M.P., Greenbaum, N.L., and Strouse, G.F. (2006) NSET molecular beacon analysis of hammerhead RNA substrate binding and catalysis. Nano Letters, 6 (7), 1318–1324. Jennings, T.L., Singh, M.P., and Strouse, G.F. (2006) Fluorescent lifetime quenching near d ¼ 1.5 nm gold nanoparticles: probing NSET validity. Journal of the American Chemical Society, 128 (16), 5462–5467. Seelig, J., Leslie, K., Renn, A., Kuhn, S., Jacobsen, V., van de Corput, M., Wyman, C., and Sandoghdar, V. (2007) Nanoparticle-induced fluorescence lifetime modification as nanoscopic ruler: demonstration at the single molecule level. Nano Letters, 7 (3), 685–689. Singh, M.P., Jennings, T.L., and Strouse, G.F. (2009) Tracking spatial disorder in an optical ruler by time-resolved NSET. Journal of Physical Chemistry B, 113 (2), 552–558. Singh, M.P. and Strouse, G.F. (2010) Involvement of the LSPR spectral overlap for energy transfer between a dye and Au nanoparticle. Journal of the American Chemical Society, 132 (27), 9383–9391. j159 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 5 How to Apply FRET: From Experimental Design to Data Analysis 84 Yun, C.S., Javier, A., Jennings, T., Fisher, 85 86 87 88 89 90 91 92 93 M., Hira, S., Peterson, S., Hopkins, B., Reich, N.O., and Strouse, G.F. (2005) Nanometal surface energy transfer in optical rulers, breaking the FRET barrier. Journal of the American Chemical Society, 127 (9), 3115–3119. Gueroui, Z. and Libchaber, A. (2004) Single-molecule measurements of goldquenched quantum dots. Physical Review Letters, 93 (16), 166108. Ko, S.H., Du, K., and Liddle, J.A. (2013) Quantum-dot fluorescence lifetime engineering with DNA origami constructs. Angewandte Chemie, International Edition, 52 (4), 1193–1197. Li, M., Cushing, S.K., Wang, Q., Shi, X., Hornak, L.A., Hong, Z., and Wu, N. (2011) Size-dependent energy transfer between CdSe/ZnS quantum dots and gold nanoparticles. Journal of Physical Chemistry Letters, 2 (17), 2125–2129. Pons, T., Medintz, I.L., Sapsford, K.E., Higashiya, S., Grimes, A.F., English, D.S., and Mattoussi, H. (2007) On the quenching of semiconductor quantum dot photoluminescence by proximal gold nanoparticles. Nano Letters, 7 (10), 3157–3164. Zhang, J., Badugu, R., and Lakowicz, J.R. (2008) Fluorescence quenching of CdTe nanocrystals by bound gold nanoparticles in aqueous solution. Plasmonics, 3 (1), 3–11. Zhang, X., Marocico, C.A., Lunz, M., Gerard, V.A., Gun’ko, Y.K., Lesnyak, V., Gaponik, N., Susha, A.S., Rogach, A.L., and Bradley, A.L. (2012) Wavelength, concentration, and distance dependence of nonradiative energy transfer to a plane of gold nanoparticles. ACS Nano, 6 (10), 9283–9290. Saraswat, S., Desireddy, A., Zheng, D., Guo, L., Lu, H.P., Bigioni, T.P., and Isailovic, D. (2011) Energy transfer from fluorescent proteins to metal nanoparticles. Journal of Physical Chemistry C, 115 (35), 17587–17593. Dexter, D.L. (1953) A theory of sensitized luminescence in solids. Journal of Chemical Physics, 21 (5), 836–850. Marcus, R.A. (1956) On the theory of oxidation–reduction reactions involving 94 95 96 97 98 99 100 101 102 103 104 electron transfer: I. The Journal of Chemical Physics, 24 (5), 966–978. Marcus, R.A. (1956) Electrostatic free energy and other properties of states having nonequilibrium polarization: I. The Journal of Chemical Physics, 24 (5), 979–989. Marcus, R.A. (1964) Chemical and electrochemical electron-transfer theory. Annual Review of Physical Chemistry, 15, 155–196. Tavernier, H.L. and Fayer, M.D. (2000) Distance dependence of electron transfer in DNA: the role of the reorganization energy and free energy. Journal of Physical Chemistry B, 104 (48), 11541–11550. Gray, H.B. and Winkler, J.R. (2005) Longrange electron transfer. Proceedings of the National Academy of Sciences of the United States of America, 102 (10), 3534–3539. Boxer, S.G. (1990) Mechanisms of longdistance electron-transfer in proteins: lessons from photosynthetic reaction centers. Annual Review of Biophysics and Biophysical Chemistry, 19, 267–299. Barbara, P.F., Meyer, T.J., and Ratner, M.A. (1996) Contemporary issues in electron transfer research. Journal of Physical Chemistry, 100 (31), 13148–13168. Benniston, A.C. and Harriman, A. (2006) Charge on the move: how electrontransfer dynamics depend on molecular conformation. Chemical Society Reviews, 35 (2), 169–179. Burshtein, A.I. (2000) Unified theory of photochemical charge separation. Advances in Chemical Physics, 114, 419–587. Farver, O. and Pecht, I. (2011) Electron transfer in blue copper proteins. Coordination Chemistry Reviews, 255 (7–8), 757–773. Fukuzumi, S., Honda, T., and Kojima, T. (2012) Structures and photoinduced electron transfer of protonated complexes of porphyrins and metallophthalocyanines. Coordination Chemistry Reviews, 256 (21–22), 2488–2502. Heckmann, A. and Lambert, C. (2012) Organic mixed-valence compounds: a playground for electrons and holes. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 160 105 106 107 108 109 110 111 112 113 114 Angewandte Chemie, International Edition, 51 (2), 326–392. Jeuken, L.J.C. (2003) Conformational reorganisation in interfacial protein electron transfer. Biochimica et Biophysica Acta, 1604 (2), 67–76. Marcus, R.A. (2006) Enzymatic catalysis and transfers in solution: I. Theory and computations, a unified view. Journal of Chemical Physics, 125 (19), 194504. Marcus, R.A. and Sutin, N. (1985) Electron transfers in chemistry and biology. Biochimica et Biophysica Acta, 811 (3), 265–322. Mayer, J.M., Rhile, I.J., Larsen, F.B., Mader, E.A., Markle, T.F., and DiPasquale, A.G. (2006) Models for proton-coupled electron transfer in photosystem II. Photosynthesis Research, 87 (1), 3–20. Moser, C.C., Anderson, J.L.R., and Dutton, P.L. (2010) Guidelines for tunneling in enzymes. Biochimica et Biophysica Acta, 1797 (8), 1573–1586. Adams, D.M., Brus, L., Chidsey, C.E.D., Creager, S., Creutz, C., Kagan, C.R., Kamat, P.V., Lieberman, M., Lindsay, S., Marcus, R.A., Metzger, R.M., MichelBeyerle, M.E., Miller, J.R., Newton, M.D., Rolison, D.R., Sankey, O., Schanze, K.S., Yardley, J., and Zhu, X.Y. (2003) Charge transfer on the nanoscale: current status. Journal of Physical Chemistry B, 107 (28), 6668–6697. Bagchi, B. (1989) Dynamics of solvation and charge-transfer reactions in dipolar liquids. Annual Review of Physical Chemistry, 40, 115–141. Bredas, J.L., Beljonne, D., Coropceanu, V., and Cornil, J. (2004) Charge-transfer and energy-transfer processes in piconjugated oligomers and polymers: a molecular picture. Chemical Reviews, 104 (11), 4971–5003. Grabowski, Z.R., Rotkiewicz, K., and Rettig, W. (2003) Structural changes accompanying intramolecular electron transfer: focus on twisted intramolecular charge-transfer states and structures. Chemical Reviews, 103 (10), 3899–4031. Kurreck, H. and Huber, M. (1995) Model reactions for photosynthesisphotoinduced charge and energy-transfer between covalently-linked porphyrin and 115 116 117 118 119 120 121 122 123 quinone units. Angewandte Chemie, International Edition in English, 34 (8), 849–866. Schuster, G.B. (2000) Long-range charge transfer in DNA: transient structural distortions control the distance dependence. Accounts of Chemical Research, 33 (4), 253–260. Algar, W.R. and Krull, U.J. (2010) New opportunities in multiplexed optical bioanalyses using quantum dots and donor–acceptor interactions. Analytical and Bioanalytical Chemistry, 398 (6), 2439–2449. Algar, W.R., Tavares, A.J., and Krull, U.J. (2010) Beyond labels: a review of the application of quantum dots as integrated components of assays, bioprobes, and biosensors utilizing optical transduction. Analytica Chimica Acta, 673 (1), 1–25. Callan, J.F., Mulrooney, R.C., Kamila, S., and McCaughan, B. (2008) Anion sensing with luminescent quantum dots: a modular approach based on the photoinduced electron transfer (PET) mechanism. Journal of Fluorescence, 18 (2), 527–532. Fan, C.H., Plaxco, K.W., and Heeger, A.J. (2005) Biosensors based on bindingmodulated donor–acceptor distances. Trends in Biotechnology, 23 (4), 186–192. Sandros, M.G., Gao, D., and Benson, D.E. (2005) A modular nanoparticlebased system for reagentless small molecule biosensing. Journal of the American Chemical Society, 127 (35), 12198–12199. Stewart, M.H., Huston, A.L., Scott, A.M., Efros, A.L., Melinger, J.S., Gemmill, K.B., Trammell, S.A., Blanco-Canosa, J.B., Dawson, P.E., and Medintz, I.L. (2012) Complex Forster energy transfer interactions between semiconductor quantum dots and a redox-active osmium assembly. ACS Nano, 6 (6), 5330–5347. Yildiz, I., Tomasulo, M., and Raymo, F.M. (2006) A mechanism to signal receptor– substrate interactions with luminescent quantum dots. Proceedings of the National Academy of Sciences of the United States of America, 103 (31), 11457–11460. Yildiz, I., Tomasulo, M., and Raymo, F.M. (2008) Electron and energy transfer j161 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 5 How to Apply FRET: From Experimental Design to Data Analysis 124 125 126 127 128 129 130 131 132 mechanisms to switch the luminescence of semiconductor quantum dots. Journal of Materials Chemistry, 18 (46), 5577–5584. Medintz, I.L., Farrell, D., Susumu, K., Trammell, S.A., Deschamps, J.R., Brunel, F.M., Dawson, P.E., and Mattoussi, H. (2009) Multiplex chargetransfer interactions between quantum dots and peptide-bridged ruthenium complexes. Analytical Chemistry, 81 (12), 4831–4839. Aslan, K., Lakowicz, J.R., and Geddes, C.D. (2005) Plasmon light scattering in biology and medicine: new sensing approaches, visions and perspectives. Current Opinion in Chemical Biology, 9 (5), 538–544. Special Issue on Plasmonics (2004). Journal of Fluorescence, 14 (4). Jain, P.K. and El-Sayed, M.A. (2008) Surface plasmon coupling and its universal size scaling in metal nanostructures of complex geometry: elongated particle pairs and nanosphere trimers. Journal of Physical Chemistry C, 112 (13), 4954–4960. Jain, P.K., Huang, W.Y., and El-Sayed, M.A. (2007) On the universal scaling behavior of the distance decay of plasmon coupling in metal nanoparticle pairs: a plasmon ruler equation. Nano Letters, 7 (7), 2080–2088. Quinten, M. (1998) Optical response of aggregates of clusters with different dielectric functions. Applied Physics B: Lasers and Optics, 67 (1), 101–106. Reinhard, B.M., Siu, M., Agarwal, H., Alivisatos, A.P., and Liphardt, J. (2005) Calibration of dynamic molecular rule based on plasmon coupling between gold nanoparticles. Nano Letters, 5 (11), 2246–2252. S€ onnichsen, C., Reinhard, B.M., Liphardt, J., and Alivisatos, A.P. (2005) A molecular ruler based on plasmon coupling of single gold and silver nanoparticles. Nature Biotechnology, 23 (6), 741–745. Su, K.H., Wei, Q.H., Zhang, X., Mock, J.J., Smith, D.R., and Schultz, S. (2003) Interparticle coupling effects on plasmon resonances of nanogold particles. Nano Letters, 3 (8), 1087–1090. 133 Tabor, C., Murali, R., Mahmoud, M., and 134 135 136 137 138 139 140 El-Sayed, M.A. (2009) On the use of plasmonic nanoparticle pairs as a plasmon ruler: the dependence of the near-field dipole plasmon coupling on nanoparticle size and shape. Journal of Physical Chemistry A, 113 (10), 1946–1953. Ullman, E.F., Kirakossian, H., Singh, S., Wu, Z.P., Irvin, B.R., Pease, J.S., Switchenko, A.C., Irvine, J.D., Dafforn, A., Skold, C.N., and Wagner, D.B. (1994) Luminescent oxygen channeling immunoassay: measurement of particle binding-kinetics by chemiluminescence. Proceedings of the National Academy of Sciences of the United States of America, 91 (12), 5426–5430. Ullman, E.F., Kirakossian, H., Switchenko, A.C., Ishkanian, J., Ericson, M., Wartchow, C.A., Pirio, M., Pease, J., Irvin, B.R., Singh, S., Singh, R., Patel, R., Dafforn, A., Davalian, D., Skold, C., Kurn, N., and Wagner, D.B. (1996) Luminescent oxygen channeling assay (LOCI(TM)): sensitive, broadly applicable homogeneous immunoassay method. Clinical Chemistry, 42 (9), 1518–1526. Elmer, P. (2013) http://www.perkinelmer. com/FR/Catalog/Category/ID/AlphaTech, March 2013. Eglen, R.M., Reisine, T., Roby, P., Rouleau, N., Illy, C., Bosse, R., and Bielefeld, M. (2008) The use of AlphaScreen technology in HTS: current status. Current Chemical Genomics, 1, 2–10. Leister, K.P., Huang, R., Goodwin, B.L., Chen, A., Austin, C.P., and Xia, M. (2011) Two high throughput screen assays for measurement of TNF-alpha in THP-1 cells. Current Chemical Genomics, 5, 21–29. Mechaly, A., Cohen, N., Weiss, S., and Zahavy, E. (2013) A novel homogeneous immunoassay for anthrax detection based on the AlphaLISA method: detection of B. anthracis spores and protective antigen (PA) in complex samples. Analytical and Bioanalytical Chemistry, 405 (12), 3965– 3972. Waller, H., Chatterji, U., Gallay, P., Parkinson, T., and Targett-Adams, P. (2010) The use of AlphaLISA technology Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 162 to detect interaction between hepatitis C virus-encoded NS5A and cyclophilin A. Journal of Virological Methods, 165 (2), 202–210. 141 Zhang, Y., Huang, B., Zhang, J., Wang, K., and Jin, J. (2012) Development of a homogeneous immunoassay based on the AlphaLISA method for the detection of chloramphenicol in milk, honey and eggs. Journal of the Science of Food and Agriculture, 92 (9), 1944–1947. j163 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Kim E. Sapsford, Bridget Wildt, Angela Mariani, Andrew B. Yeatts, and Igor Medintz 6.1 Introduction The intrinsic sensitivity (r6 dependence) of F€orster (or fluorescence) resonance energy transfer (FRET) to nanoscale changes in the donor/acceptor separation distance has made FRET an invaluable biophysical tool in a variety of applications, ranging from studying the structure and conformation of proteins and nucleic acids to examining biomolecular interactions, including its use in in vitro and in vivo bioassays [1–7]. While the myriad of FRET configurations and techniques currently in use are covered throughout this book, here we focus primarily on the materials utilized as donor or acceptor probes in FRET rather than the process itself [3,5,8,9]. Our 2006 review paper on this topic serves as the foundation for this updated chapter [3]. The materials were divided into three main categories: organic materials that include “traditional” dye fluorophores, dark quenchers, polymers, and carbon nanomaterials (NMs); inorganic materials such as metal chelates, metal, and semiconductor nanocrystals; and fluorophores of biological origin such as fluorescent proteins (FPs), amino acids, and fluorescence generated from enzymatic bioluminescence (BL) and chemiluminescence (CL). These materials may function as FRET donors and/or acceptors, depending upon experimental design. Many of the new materials developed and/or new donor–acceptor probe combinations used address some of the inherent complications of more traditional FRET materials, including photobleaching, spectral cross talk, and direct excitation of the acceptor species, and examples of these will be highlighted and discussed throughout the chapter. Since the vast majority of FRET applications are biological in nature, they routinely involve some type of biomolecule labeling strategy, which ultimately plays a significant and fundamental role in the success and interpretation of the resulting FRET. Therefore, we begin the chapter with a brief discussion of the bioconjugation techniques commonly utilized for FRET and points to consider, followed by sections highlighting the current and emerging materials that are used, or have the potential to be used, in FRET applications. FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j165 j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 6.2 Bioconjugation A large number of FRET-based applications involve the use and labeling of some type of biomolecule (e.g., cell membrane, antibodies, nucleic acids, protein, and peptides). Given the sensitivity of FRET to a number of parameters, the ability to control the donor/acceptor labeling in the system under investigation is paramount to obtaining well-controlled and reproducible results that will aid in subsequent interpretation of experimental data. The extent of such control is initially dictated by a combination of factors, including (i) the nature of the system under study (i.e., are inter- or intramolecular studies desired?), (ii) the availability/number (and reactivity, where applicable) of attachment/incorporation sites on the biomolecule(s) for the donor/acceptor probes, (iii) the nature of the donor/acceptor probes (e.g., organic molecule, fluorescent protein, and NM), (iv) the size of the donor/acceptor probes (especially protein-based and NMs) relative to the biomolecule(s), which can influence the system under study, and (v) the availability of the donor/acceptor probes with the desired reactivity for bioconjugation and the nature of the linker (e.g., length and flexibility), connecting the donor/acceptor probes to the biomolecule. For quantitative FRET the microenvironment dependency of the fluorophore’s photophysical properties and the uncertainty in probe position and orientation relative to the biomolecule should be considered when choosing donor/acceptor probes and bioconjugation strategies [6]. Donor/acceptor probe attachment to a biomolecule can be achieved via a number of labeling techniques that can be chemically or biologically inspired in nature (Figure 6.1); for recent reviews see Refs [10–12]. Bioconjugation based on forming interactions, typically covalent bonds between the biomolecule and the probes, represents the most popular and traditional group of chemistries used to date, and will likely remain the workhorse in the near future due to the commercial availability of a wide variety of donor/acceptor probes modified with a number of reactive functionalities that facilitate bioconjugation. In the case of protein labeling, for example, the predominant chemically based bioconjugation strategies target the naturally occurring amino acids lysine (Lys – primary amine) and cysteine (Cys – thiol) with succinimidyl ester (NHS) or maleimide reactive groups, respectively. Control of the donor/acceptor probe locations and stoichiometry is one of the many important considerations when designing FRET studies [5]. In the case of deoxyribonucleic acid (DNA)/ribonucleic acid (RNA) and short synthetic peptides, control of location and stoichiometry can be programmed into their structure via inclusion of the donor/acceptor molecules themselves or site-specific incorporation of thiol/amine groups for subsequent labeling during synthesis [13,14]. However, proteins are generally more complex because they contain a number of primary amines that can cause difficulty in controlling the location of labeling and usually results in variable dye-to-protein (D/P) ratios. Targeting thiols on Cys residues with maleimide chemistry is more specific and can reduce the labeling variation, as these residues are far rarer. However, Cys residues are often critical to protein structural conformations as Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 166 Figure 6.1 Bioconjugation methods separated into chemistry- and biology-inspired techniques. part of disulfide bonds and are typically buried below the protein surface, which can limit access depending on the size of the chosen donor/acceptor probe. Thiols can be chemically (typically via Lys interconversion) or recombinantly introduced into the protein surface [15–17]. However, this too can be problematic as additional Cys residues can “thiol-scramble” the protein structure during folding, and surfaceexposed thiols can result in the formation of protein dimers, trimers, and so on, which, when purified, necessitate further reduction prior to labeling. There are a couple of excellent resources available for researchers interested in bioconjugation protocols in general [18] and fluorescent labeling in particular [19]. Bioorthogonal reactions (reactions that do not interfere with other biogroups besides the target), which were born out of the desire to study biomolecules in their native environment, have the ability to address many of the controlled labeling concerns outlined earlier [12,20]. Many of these reactions are based on organic chemistry-inspired covalent modifications and include the Staudinger ligation j167 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.2 Bioconjugation j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations reaction, ketone/aldehyde condensation reactions (bioorthogonal, depending on the system under study), and a variety of cycloaddition reactions, including the quintessential “click chemistry” azide/alkyne cycloaddition. There are a number of biologically inspired bioconjugation strategies that are gaining popularity, which, although perhaps not bioorthogonal in the strictest sense, possess many of the stipulations and benefits required to be considered as bioorthogonal reactions [12,21]. Fluorescent proteins (FP) such as green fluorescent protein (GFP) can be appended to existing proteins using recombinant techniques generating protein chimeras, and although not considered a bioorthogonal chemistry, this does allow the site-specific incorporation of the FP to the protein of interest (see Section 6.4.3) [22]. Likewise, nonnatural amino acid residues (Section 6.4.2), short peptide tags, and full proteins (such as enzymes), that are either fluorescent or allow specific labeling with fluorophores, can also be incorporated genetically into protein structures [10,12,23]. For example, the tetracysteine/biarsenical system, originally developed by Tsien and coworkers, demonstrated that proteins expressing an optimized Cys-Cys-X-X-Cys-Cys sequence [where X ¼ could be any amino acid, but is traditionally proline-glycine (Pro-Gly)] would react with biarsenical-functionalized fluorophores (e.g., FlAsH and ReAsH) (Figure 6.2) [21,24,25]. Oligohistidine (His) peptides (e.g., His6) are peptide tags that have been used for conjugation. They are known to bind nickel-nitrilotriacetic acid (Ni2þ-NTA)-functionalized molecules, while the oligoaspartate (Asp) sequence (D4-tag, Asp4) has a strong binding affinity to multinuclear zinc(II) complexes [10,12,26,27]. There are an expanding number of biologically inspired techniques that harness genetically encoded peptide handles combined with enzymes to catalyze small-molecule (e.g., biotin and fluorescent dye) conjugation [10–12,16,23,28,29]. Figure 6.2 Bioconjugation using peptide recognition. (a) Schematic showing the labeling of an engineered protein displaying a linear tetracysteine motif with bisarsenical containing fluorophores. (b) Chemical structures of FlAsH and ReAsH. (Reprinted with permission from Ref. [25]. Copyright 2011, American Chemical Society.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 168 Jager et al., for example, modified a model protein, chymotrypsin inhibitor 2 (CI2), with a transglutaminase (TGase)-tag (Pro-Lys-Pro-Gln-Gln-Phe, where Gln is glutamine and Phe is phenylalanine) at its N-terminus [16]. TGase facilitated the coupling of the TGase-tag-modified CI2 with AlexaFluor 647 (Alexa647, A647) cadaverine (A647-(CH2)6-NH2), forming an isopeptide bond between the TGase-tag Gln residue and the primary amine on the A647 fluorophore. In another example, the Escherichia coli enzyme biotin ligase has been shown to ligate biotin to proteins tagged with an acceptor peptide sequence Lys-Lys-Lys-Gly-Pro-Gly-Gly-Leu-Asn-Asp-Ile-Phe-GluAla-Gln-Lys-Ile-Glu-Trp-His (where Leu is leucine, Asn is asparagine, Asp is aspartic acid, Ile is isoleucine, Ala is alanine, and Trp is tryptophan) [28]. The biotin ligase also accepted a ketone isostere of biotin as a cofactor resulting in ketone-functionalized proteins that could be subsequently modified with hydrazide- or hydroxylamine-functionalized molecules. The HaloTagTM [haloalkane dehalogenase (DhaA)] and SNAPTM-tag [O6-alkylguanine-DNA alkyltransferase (hAGT)] are labeling techniques that actually involve fusing the full enzymes (which self-label themselves) to the protein of interest. The HaloTag utilizes fluorescently labeled haloalkane substrates, while the SNAP-tag uses fluorescent benzylguanine derivatives to generate fluorescently labeled targets [10–12]. Nanomaterials are increasingly being used as donor/acceptor probes in FRET studies due to their many unique properties, and as such many of the bioconjugation techniques described earlier are applicable here as well [8,9]. However, taking into account their nanoscaffold nature, additional concerns should be considered during bioconjugation [5,30–32]. NM surfaces can be quite complex, comprising not only the NM itself but also oftentimes additional stabilizing ligands that help maintain its aqueous solubility (e.g., colloidal stability) and prevent undesirable interactions such as aggregation–agglomeration. In addition NMs are generally much larger than your typical fluorescent/quencher organic molecules and are often on a similar size and scale or larger than most biomolecules, therefore the potential influence of this size on the system under study should be carefully considered. Other factors to consider regarding bioconjugation include how the NM is stabilized in solution and whether the reaction conditions might affect the stability or physical properties of the NM. A prime example that highlights this issue is the use of carbodiimide coupling chemistry to link carboxylic acid (COOH)-modified NMs with the primary amines on a protein, as discussed in a recent review by Algar et al. [30]. COOH-terminated ligands are commonly used to impart charge-based solubility and stability to inorganic NMs at basic pH. However, carbodiimide activation (often added in huge excess due to rapid hydrolysis of the reaction intermediate) converts the COOH to a less soluble o-acylisourea intermediate and hence can cause reduced solubility and NM aggregation, severely impacting product quality and reproducibility. Another concern is whether the stabilizing surface ligands influence subsequent interactions involving either the biomolecule attachment reaction or the subsequent biorecognition event. Dennis et al. used FRET to investigate eight distinct quantum dot (QD) coatings and their influence on the self-assembly of a His-tag-labeled mCherry FP that interacts directly with the QD surface via metal affinity coordination [33]. This coordination relies substantially on access of the j169 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.2 Bioconjugation j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations His-tag to the QD surface, and the researchers found that even subtle changes in the organic coating could significantly alter the accessibility and hence His-tag– mCherry coordination. In another example of reaction conditions that can impact the NM, the copper (Cu) in the widely used Cu-catalyzed click chemistry has been found to be extremely detrimental to the luminescent properties of QD materials [34]. As an alternative Bernardin et al. used a Cu-free strained click chemistry technique, coupling strained cyclooctyne-functionalized QDs with azido biomolecules, and the QDs in this case retained their strong luminescent properties [34]. Surface ligands and the attachment chemistry used in bioconjugation can also affect the performance of the NM bioconjugate in the desired application. For example, while developing a FRET-based QD–peptide sensor for monitoring botulinum neurotoxin A (BoNT A) activity, Sapsford et al. found that the stabilizing ligand sterically hindered the BoNT A from interacting with the peptide substrate assembled on the QD surface via metal affinity coordination, resulting in a nonresponsive sensor [35]. The situation was mitigated by conjugating the peptide to the terminal groups of the QD-stabilizing ligand, improving access for the BoNT A. As with any type of bioconjugation involving a surface, the conformation/orientation of the biomolecule upon immobilization is an essential component of its subsequent functionality and should be considered during study design. Loss of biomolecular activity is to be expected if the recognition site of the biomolecule is positioned in close proximity to the surface [30]. In a study of QD–DNA bioconjugates, Boeneman et al. found that the attachment chemistry strongly influenced the orientation of DNA on a QD-poly(ethylene glycol) (PEG) surface [36]. His-tag-modified DNA attached directly to the QD surface resulted in a structure that, as predicted, extended out from the surface. Biotin-labeled DNA bound to streptavidin (SA)modified QDs, however, did not follow predicted models, and the DNA was found to take a number of random orientations on the QD surface, which was attributed to the random attachment of SA to the QD. Random orientations are likely to result in a distribution of biomolecular activities, which can affect reproducibility of experimental results [30]. Clearly, there is a wide range of bioconjugation techniques available to researchers, and choosing the most suitable method depends on the nature of the system under study. Some of these labeling techniques can be complex, requiring for example genetic engineering expertise; however, the increase in commercial reagents and kits (from companies such as Life TechnologiesTM, Sigma-Aldrich, and Promega) for these types of bioconjugation reactions, especially with the introduction of newer bioorthogonal chemistries, is encouraging more widespread adoption. Regardless of the method chosen, having both the donor/acceptor probes at known and distinct locations on biomolecule(s) is most desirable in terms of postexperimental analysis. Care should be taken to ensure that donor/acceptor probe modification does not influence the biomolecule functionality, especially when trying to determine the functional characteristics of a biomolecule (where labeling may interfere with/alter structure, biomolecule conformational changes, or biomolecule interactions). Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 170 6.3 Organic Materials As highlighted in this section, organic materials that possess the necessary photophysical properties to be utilized as FRET donors or acceptors are a widely diverse group and include molecules, macromolecules, polymers, and NMs. 6.3.1 Ultraviolet, Visible, and Near-Infrared Emitting Dyes The majority of donor/acceptor materials currently used in FRET applications are ultraviolet (UV), visible (Vis), and near-infrared (IR) emitting organic dyes. These “traditional” dyes are usually the first type of FRET material tested with other new or non-“traditional” fluorescent materials in potential FRET systems. The most common organic dye classes, shown in Figure 6.3, have several advantages, such as commercial availability, cost-effectiveness, extensive characterization of FRET properties, easy bioconjugation through NHS-ester, maleimide, hydrazide, or amine chemistries, availability in reactive form, and high quantum yields (QYs) and solubilities [37,38]. Recent advances in click chemistry now allow companies such as Lumiprobe to provide organic dyes conjugated to azide and alkyne chemical moieties for further bioconjugation as well. Despite all the advantages of these traditional dyes, there are disadvantages too. For instance, some have a high rate of photobleaching, may be sensitive to pH, and have a propensity to self-quench when highly substituted on biomolecules. Some of the redder dyes have low solubility in aqueous solvents and for FRET in particular, the broad absorption/emission profiles and small Stokes shifts often lead to direct excitation of the acceptor, complicating subsequent analysis. The quest for new organic dyes with the potential to overcome these limitations continues, most recently with materials such as the Chromeo [39] (currently sold by Active Motif), CS1-6 near-IR (NIR) [40], and alkyne carbocation “cyanine-like” dye families [41–43]. Researchers have recently demonstrated a series of organic molecules that undergo excited-state intramolecular protein transfer (ESIPT) for use in the development of fluorescent chromophores with a large Stokes shift (LSS) [44–46]. Many resources are available to aid in choosing suitable donor–acceptor pairs, including a number of FRET reviews [15,47–49], as well as the Molecular Probes Handbook [19] and a review by Wu and Brand [50] that offers an extensive list of donor–acceptor dye pairs and their respective R0 values. Life Technologies’ Fluorescence SpectraViewer (http://www.invitrogen.com/site/us/en/home/support/ ResearchTools/Fluorescence-SpectraViewer.html) and Zeiss’ Fluorescence Dye and Filter Database (https://www.micro-shop.zeiss.com/us/us_en/spektral.php) are Web-based programs that allow researchers to plot multiple dye absorption/ emission profiles to optimize spectral overlap and choose appropriate filters. An excellent comparison of the physical and spectroscopic properties of a number of red-absorbing dyes is provided by Buschmann [51]. See also the extended and updated tables collated by van der Meer and provided later in this book. j171 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3 Organic Materials Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 172 Many current applications still rely on traditional dye–dye FRETcombinations due to the many advantages of organic dyes as described earlier [3,47,52]. For instance, FRET-based biosensors have led to a deeper understanding of a number of biological phenomena such as integrin adhesiveness and signaling dynamics [53], plasma membrane biophysical interactions [54], host–pathogen interactions [55], and protein folding geometries and conformational states [56]. Similar dye combinations are also useful for FRET-based biosensing, including glucose sensors [57] and biological agent detection [58]. Dye–dye FRET combinations have had tremendous impact on biomedical research, specifically in the areas of nucleic acid analysis, DNA sequencing, and genotyping [13,59–62]. In addition, molecular beacon probes used in nucleic acid analysis and Scorpion real-time PCR assays are often FRET based [13,63]. Interestingly, use of DNA scaffolds incorporating donor/acceptor dyes has led to a more fundamental understanding of the orientational dependence of the dyes on FRET efficiency [64–66]. DNA microarrays, in which FRET-based DNA probes are immobilized on solid surfaces, are quickly becoming an exciting application with the potential to increase sensitivity, specificity, and throughput of gene expression as well as large-scale single-nucleotide polymorphism (SNP) discovery, detection, and genotyping [67]. There is no doubt that new applications will continue to drive the development of novel donor–acceptor dye combinations that overcome current deficiencies in existing organic dyes. 6.3.2 Quencher Molecules The use of quenching molecules in FRET-based applications continues to be popular. The primary advantage of using these molecular acceptors over their fluorescent counterparts is the elimination of background fluorescence due to direct acceptor excitation or reemission. Typically, quenchers take the form of organic molecules or metallic materials such as gold (Au) (Section 6.5.3). Figure 6.4 offers a visual representation of a variety of organic quencher families that are commercially available. Two of the most common quenching acceptor molecules, Dabcyl (4,-(40 -dimethylaminophenylazo)benzoic acid) and Dabsyl (4-dimethylaminoazobenzene-40 -sulfonyl), have absorption maxima centered at 485 and 466 nm, respectively. Another recent addition to nonfluorescent quencher dyes is IRDye QC-1, which is characterized by a broad absorption peak between 550 and 950 nm, 3 Figure 6.3 Organic UV and visible fluorescent dyes. (a) Structures of the common organic UV and visible fluorescent dyes. Typical substituents at the R position include CO2, SO3, OH, OCH3, CH3, and NO2; Rx marks typical location of the bioconjugation linker. (Reprinted with permission from Ref. [3].) (b) Plot of fluorophore brightness versus the wavelength of maximum absorption (max) for the major classes of fluorophores. The color of the structure indicates its wavelength of maximum emission (em). For clarity, only the fluorophoric moiety of some molecules is shown. (Reprinted with permission from Ref. [37]. Copyright 2008, American Chemical Society.) j173 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3 Organic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.4 Organic quencher molecules. (a) Example of structures of the common quencher molecules. Substituents R are listed. Rx marks typical location of the bioconjugation linker. (Reprinted with permission from Ref. [3].) (b) Some common, commercially available, quencher families along with absorbance maxima and spectral regions covered by a particular quencher family. (Adapted with permission from Ref. [3].) effectively allowing the quenching of both near-infrared and the more commonly used visible donor dyes [68]. Other quencher families that tend to have wide-range absorption spectra include the trademarked QSY, QXL, ATTO, BlackBerry, and Black Hole Quenchers. These broad absorption spectra decrease design constraints and allow quencher molecules to function as acceptors for many dyes. One application where quenchers are often applied is DNA analysis, specifically molecular beacons coupled with organic dye donors [13,48,49]. The primary advantage of this donor– Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 174 quencher configuration is that singular or individual donor channels can be monitored and, if sufficient spectral separation is achieved, utilized for “multiplexing” with a variety of other donor–quencher pairs. Typical applications of these methods have measured DNA permeability of polyelectrolyte thin films [69] and utilized catalytic DNA biosensors to detect lead (Pb) ions [70]. Another technique employs quencher-labeled substrate analogues used in conjunction with dye-labeled proteins for FRET-based biosensing of nutrients through displacement [71]. Reduced size of donor/quencher pairs can also improve the accuracy in determining protein dynamics. For example, the use of thioamide as a quencher fabricated from the protein’s own backbone, through a single-atom substitution, minimized perturbation of the system [72]. Current techniques also look into increasing the efficiency of quencher molecules through the use of binding molecules with advantageous three-dimensional conformations and charge density allowing increased local dye concentrations [73]. One of the few examples of FRET, where organic quencher molecules are coupled to nonorganic fluorophores, involves QD donors (see Section 6.5.5). 6.3.3 Environmentally Sensitive Fluorophores While the vast majority of fluorophores will respond to a certain extent to a perturbation in their microenvironment, some exhibit much higher sensitivity than others and as such are classified as environmentally sensitive. These fluorophores exhibit some change in their spectral characteristics (absorption/emission profiles) in response to a change in their microenvironment, such as pH, ion interactions, or another moiety such as oxygen (O2), solvation, polarity (solvatochromic fluorophores), or rigidity, and these dyes are usually defined by the analyte or condition that they respond to most favorably [19,47]. A large number of environmentally sensitive fluorophores are found in the small organic molecules class of fluorescent probes, however, other classes of fluorophores such as FPs have also been found to be sensitive to specific changes in their environment [37,74–79]. Life Technologies offers a wide range of environmentally sensitive fluorophores, including dyes sensitive to reactive oxygen species (ROS), pH, calcium (Ca2þ), magnesium (Mg2þ), Zn2þ, sodium (Naþ), chloride (Cl), potassium (Kþ), and membrane potential [19]. The dyes are offered in a variety of forms depending on the application, including cell-permeant and cell-impermeant, or can be modified with functional groups that aid in subsequent conjugation if desired. Many of the target ions are important signaling molecules in molecular cell biology, and intracellular measurements are essential to understanding the processes that govern cellular function [76,77]. In the case of intracellular pH, for example, two functional microenvironments should be considered, the cytosol (pH 6.8–7.4) and the acidic organelles (pH 4.5–6.0), with the exact choice of fluorescent probe dependent on its pKa [19,76]. Oregon Green and LysoSensors are appropriate for the more acidic organelle environment, while fluorescein derivatives, the polar 20 ,70 -bis-(2-carboxyethyl)-5-(and-6)-carboxy-fluorescein (BCECF) derivative (developed by Tsien) and j175 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3 Organic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations the proprietary seminaphthorhodafluors (SNARF) dyes, function optimally in the pH 6.0–8.0 range and are popular choices for cytosol pH measurements [19,80]. Nakata et al. recently developed two new SNARF derivatives, SNARF-F and SNARFCl, which, while maintaining the characteristic spectral changes of the original SNARF dye, have improved cell permeability and in the case of SNARF-F an improved pKa (7.38) for cytosol measurements (Figure 6.5) [81]. Tweaking of the fluorophore chemical structure can improve the fluorescent properties, and the target interactions of many of these fluorescent probes and combinatorial approaches offer an interesting alternative to more traditional and rational design methods for identifying new improved fluorescent candidates, recently reviewed in Ref. [82]. While environmentally sensitive fluorophores function alone adequately as qualitative intensity-based fluorescent probes, the combination of FRET using environmentally sensitive dyes has been demonstrated for detection of pH, ammonia (NH3), and carbon dioxide (CO2) [83,84]. When target quantification is desired, especially in studies involving complex cellular environments, FRET represents one mechanism in which to achieve ratiometric measurements that are independent of fluorescent signal fluctuations [77,85]. Signal fluctuations can result from variations in local probe concentration, sample thickness, pH, or temperature, which make Figure 6.5 Environmentally pH-sensitive SNARF dyes. (a–c) Absorbance (solid line) and fluorescence (dashed line) spectra of 10 mM of (a) SNARF, (b) SNARF-F, and (c) SNARF-Cl at pH 5.0 (red) and pH 10.0 (blue), excited at the isosbestic point, respectively. (d and e) Photos of (d) color and (e) fluorescence of SNARF (10 mM), SNARF-F (10 mM), and SNARF-Cl (10 mM) at pH 5.0 and pH 10.0. (Reprinted with permission from Ref. [81]. Copyright 2011, Elsevier.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 176 subsequent interpretation of single-intensity data complicated. Ratiometric FRETbased sensors for the analysis of nucleoside polyphosphates, pH, temperature, hydrogen peroxide (H2O2), mercury (Hg), and chromium (Cr) ions have all been demonstrated [79,86–90]. Environmentally sensitive fluorophores are commonly incorporated into nanoparticles (NPs) for sensing applications. For example, Childress et al. developed dyedoped polymer NPs for ratiometric fluorescence detection of Hg(II) ions in aqueous solution [89]. NPs of the conjugated polymer (CP) poly[(9,9-dioctylfluorenyl-2,7diyl)-co-(1,4-benzo-{2,10 -3}-thiadiazole)] (PFBT), which fluoresced green-yellow, were doped with a nonfluorescent Hg(II)-responsive rhodamine dye. The rhodamine dye, upon interacting with Hg(II) ions, converts to an orange-red fluorescent form, resulting in FRET between the donor PFBT NPs and the activated acceptor rhodamine dye. The sensing NPs could measure low levels of Hg(II) ions in the 0.7–10 ppb range, and the ratiometric nature of the FRET sensor eliminated any issues related to environmental or instrumental fluctuations. Using a more conventional platform, Kurishita et al. developed FRET-based ratiometric chemosensors for detection of nucleoside polyphosphates such as adenosine-50 -triphosphate (ATP) [87]. The sensor was based upon a large off–on fluorescence enhancement that occurred when a xanthene-based Zn(II) complex bound ATP (Figure 6.6a). The chemosensor combined a coumarin (blue fluorescence) donor with the xanthenebased Zn(II) complex acceptor, which developed green fluorescence upon ATP binding. Ultimately the chemosensor was tested in live Henrietta Lacks (HeLa) cells, where initial staining resulted in green fluorescence due to the presence of ATP in the cells (Figure 6.6c). Introduction of 2-deoxyglucose (2-DG) and/or potassium cyanide (KCN), which inhibited ATP synthesis, resulted in a significant decrease in FRET, causing an increase in the blue donor emission compared to untreated cells (Figure 6.6d and e). Nucleic acid binding dyes represent a specific class of environmentally sensitive organic dyes that warrant special mention [91,92]. These dyes form complexes with nucleic acids, resulting in significant off–on fluorescence enhancements, which also incorporate intermolecular homo-FRET to a large extent. Dyes include 40 ,6-diamidino-2-phenylindole (DAPI), the bisbenzimidebased dyes (collectively called the Hoechst dyes), OliGreen, ethidium bromide (EtBr), propidium iodide, and the cyanine dyes (PicoGreen, YOYO, and TOTO families of dyes; SYBR Green I and SYBR Gold) [91,92]. Cosa et al. performed a comprehensive study comparing the photophysical properties of a number of these dyes alone and upon binding single-stranded DNA (ssDNA) or doublestranded DNA (dsDNA) [91]. All these dyes interact with dsDNA, with some such as DAPI and the Hoechst dyes, binding specifically with the minor groove of adenine (A)–thymine (T)-rich sequences [91,92]. Most of the dyes also seem to interact with ssDNA and some, such as EtBr, have also been found to interact with RNA [91,92]. The combination of nucleic acid sensitive dyes and FRET has been used in a number of studies dealing with understanding how the dyes interact with DNA [92–94] and in the detection of single-nucleotide polymorphisms (SNPs) and short tandem repeats (STRs) [95–97]. j177 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3 Organic Materials Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3.4 Dye-Modified Microspheres/Nanomaterials There are a wide range of materials that can and have been modified with organic dyes to generate fluorescent microspheres or NMs, although the most common platforms are either polymer or silica based. Modification typically involves either encapsulation of the fluorophores within the material core or interaction with the particle surface. Recent advances in NM synthesis and improved surface functionalization techniques have allowed the increased use of dye-labeled NMs in particular, which, from a size perspective, is a benefit in FRET-based applications. Dye-labeled nano- and micro-sized particles have been prepared through ionic interaction [98], miniemulsion [99,100], covalent conjugation [101], or encapsulation of fluorophore molecules during synthesis [102–105], and the benefits of these formats result in increased signal, decreased photobleaching, increased surface functionalizability, and lower limits of detection than small-molecule fluorophores [106]. Due to the popularity of diagnostic and research techniques utilizing fluorescence detection systems, commercial fluorescent particles are available from a wide variety of sources such as Molecular Probes, Phosphorex, Spherotech, Bangs Laboratories, Sigma-Aldrich, and Polysciences. Besides standard fluorophores, fluorescent microspheres loaded with europium (Eu) chelates are available, extending the utility of such labels to time-resolved energy transfer configurations (Section 6.5.1) [107]. A current trend in microspheres and NM synthesis involves using FRET to tune the spectral properties of the resulting fluorescent microspheres/NMs, generating fluorescent FRET-based tags that have improved Stokes shifts and unique fluorescent signatures that can be excited by a single wavelength (Figure 6.7) [99,100,102,108–111]. These FRET systems can also be designed to contain Figure 6.7 Polymeric FRET-based NPs for in vivo imaging. The particles were assembled from diblock copolymers of poly(D,L-lactic-coglycolic acid) and maleimide-activated PEG, which were also encapsulated in both the donor (1,10 -dioctadecyl-3,3,30 ,30 tetramethylindodicarbocyanine) and acceptor (1,10 - dioctadecyl-3,3,30 ,30 tetramethylindotricarbocyanine) fluorophores. FRET resulted in a large Stokes shift (>100 nm) of the emission maxima, and the transfer efficiency could be fine-tuned by further adjusting the doping ratio of the donor and acceptor fluorophores. The optimized formulation was less than 100 nm in size, brighter than quantum dots, stable in biological media, and demonstrated similar biodistribution to most NMs. (Reprinted with permission from Ref. [102]. Copyright 2012, American Chemical Society.) j179 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3 Organic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations photochromic dyes (discussed further in Section 6.3.6) that increase signal detection by decreasing background emission [99]. A variety of hybrid fluorescent core–shell silica NPs with interesting optical properties have been developed, such as, dye-labeled silica-coated silver (Ag) NPs with enhanced FRET properties [112,113] and Cornell dots (CU dots) [114,115] developed by Wiesner at Cornell University. CU dots are synthesized by covalently conjugating dye molecules to a silica precursor before being condensed to form a dye-rich core. Silica sol–gel monomers are then added to form a denser outer silica network. The use of dye-labeled microparticles and NPs in FRET-based analytical and research assays continues to grow as these materials are exploited in nanoscopic ruler measurements [116], biosensing of infectious agents [105,111], in vivo biolabeling [102], flow cytometry [117,118], SNP genotyping [119,120], and disease diagnosis using protein nanoarrays [121]. 6.3.5 Dendrimers and Polymer Macromolecules Dendritic and polymer macromolecules are increasingly being used in fluorescence-based applications. Dendritic architectures represent a class of repeatedly branched or tree-like polymeric structures of which the subclasses of dendrimers and dendrons have found particular application in molecular imaging, sensing, photovoltaics, and energy harvesting [122–124]. Dendrimers, in particular, are highly ordered macrostructures comprised of a distinct core, branched mid, and branched surface/periphery regions. The ability to tailor dendrimer structures with inclusion of multiple functional groups during synthesis allows precise control over the position and orientation of fluorophores and attachment of biomolecules leading to well-defined macromolecules [122–124]. This precise structural control has made dendrimers interesting synthetic platforms in which to mimic Nature’s light-harvesting structures, reviewed in Ref. [124], and in the development of new photovoltaic materials [125]. Of the many types of dendrimers reported, the poly(amidoamine) (PAMAM)-based materials are the most utilized because the many primary amines enable facile functionalization [122,123]. The ability to label these dendritic structures with multiple fluorophores results in increased absorption cross sections and higher fluorescence intensities, which is ideal for bioassay and imaging applications [122,123]. This strategy is employed in a class of dendron-based fluorogenic dyes known as dyedrons, which comprise multiple donor Cy3 dyes coupled to a single malachite green (MG) acceptor [126]. Free in solution, the MG acceptor acts as a quencher due to unconstrained internal structural rotations, however, upon binding to fluorogen-activating proteins (FAPs) such as single-chain variable fragment antibodies, the MG acceptor becomes activated, making it highly fluorescent when excited via FRET from the Cy3 donors. Such dyes have potential application as targeted probes in sensitive homogeneous (no wash) imaging. Inherently fluorescent cationic dendritic structures comprising primarily phenylene-ethynylene Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 180 have been used as donors for the detection of peptide nucleic acid (PNA)/DNA hybridization, where the neutral PNA probe was labeled with a fluorescein acceptor [127]. Davis et al. used inherently fluorescent cationic and anionic diphenylacetylene dendrimers doped into cellulose acetate to generate solid-state electrospun nanofiber sensor arrays for protein detection [128]. These nanofiber films selectively interacted with proteins (including metal- and nonmetal-containing proteins), which resulted in fluorescent quenching in distinct patterns, due to varying interactions with the different protein structures; this allows specific protein identification even in complex mixtures. Dendrimers have also been used as carriers for multiple materials such as combinations of drugs, nucleic acids, antibodies, fluorescent tags, and/or contrasting agents that can be used for imaging, drug delivery, or bioassay applications [123]. For example, Myc et al. used folic acid modified dendrimers to target delivery of a FRET-based apoptotic sensor to cancer cells that could be used to determine the efficacy of a chemotherapeutic or other targeted treatment by monitoring cell death in real time [129]. Although synthetic procedures for dendrimer synthesis are available in the literature, dendrimers are also commercially sold with functionalities that can be further modified by the end user, which may aid in more widespread use (Dendritech, Polymer Factory Sweden AB, and Dendritic Nanotechnologies Inc.). Qiagen offers dendrimers specifically functionalized to bind both DNA and cells for cellular transfection. Fluorescent polymers are a related class of fluorophores that can be either intrinsically fluorescent such as CPs or functionalized with multiple fluorophores [3,130]. Similar to dendrimers, fluorescent polymers are characterized by large molar absorption coefficients and therefore high fluorescence. However, due to their inherent polydispersity the emission from fluorescent polymers is typically not localized, resulting from energy transfer processes along the whole chain, with a net result of diffuse emission [131]. Thus fluorescent polymers cannot be considered point donors for FRET. Nonetheless, fluorescent polymers, especially CPs, have found application in a number of fluorescence-based studies, including FRET [130]. Fluorescent CPs, cationic conjugated polymers (CCPs), and conjugated polyelectrolytes (CPEs) are particularly popular for developing FRET-based DNA biosensors [132], with applications ranging from detection of DNA [133,134], DNA hybridization [135], DNA methylation [136], and SNPs [137]. They have also been incorporated into sensing schemes for proteins [138] and Hg ions [89,139]. The group of McNeill and coworkers recently developed a series of multicolor conjugated polymer dots (CPdots) (Figure 6.8) [140–142]. These polymer dots, referred to as CPdots and later Pdots, are made from semiconducting polymer materials such as PPE, PFPV, or PFBT (see Figure 6.8 for full chemical name), are 4 nm in diameter with high fluorescent intensities, and could be readily functionalized [140–142]. Others have studied these Pdot materials, including Chan et al. who developed FRET-based Pdots with photoswitching capabilities [143,144]. These were created through the incorporation of photochromic spiropyran molecules into a PFBT polymer and were intended for use in bioimaging applications. In another example, a pH-sensing Pdot was created by functionalizing PPE Pdots with pH-sensitive fluorescein [145]. j181 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3 Organic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.8 Conjugated polymer dots (CPdots). (a) Chemical structures of fluorescent CPdot: the polyfluorene derivative poly(9,9dioctylfluorenyl-2,7-diyl) (PFO), the copolymer poly[{9,9-dioctyl-2,7-divinylene-fluorenylene}alt-co-{2-methoxy-5-(2-ethylhexyloxy)-1,4phenylene}] (PFPV), the poly[(9,9dioctylfluorenyl-2,7-diyl)-co-(1,4-benzo-{2,10 ,3}thiadiazole)] (PFBT), and the poly(phenylene vinylene) derivative poly[2-methoxy-5-(2ethylhexyloxy)-1,4-phenylenevinylene] (MEHPPV). (b) Photographs of aqueous CPdot suspensions under room light (left) and UV light (right) illumination. (c) Absorption spectra and (d) fluorescence spectra of the conjugated polymer dots. (Reprinted with permission from Ref. [140]. Copyright 2008, American Chemical Society.) 6.3.6 Photochromic Dyes Materials that can reversibly switch between two forms/states upon exposure to electromagnetic radiation, in which each form/state has different photophysical properties, are known as photochromic dyes [146,147]. While there are both inorganic and organic photochromic materials, organic photochromic dyes are the most popular, having a wide range of uses from decoration and eyeglass lens Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 182 coatings to optical switches and data storage. There are a number of photochromic dyes and various mechanisms that cause their photochromic transformations, and these have been extensively reviewed [146–150]. Of these, spiro-based (e.g., spiropyran) and increasingly diarylethene-based photoswitchable compounds have found particular application in FRET, so-called photochromic FRET (pcFRET) [147,149,151–153]. The pcFRET technique is particularly useful in FRET imaging applications, especially on a single-protein level, where pcFRET can be used to turn the FRET process “off” or “on” thereby creating an internal control and eliminating false-positive or false-negative signals due to high intrinsic autofluorescence, interactions with other endogenous proteins, and/ or low FRET efficiencies [152,154–157]. FRET imaging sensitivity can be further enhanced when pcFRET is used in conjunction with techniques such as optical lock-in detection (OLID) [156,158]. Spiro-based materials represent one of the more prominent types of photochromic dyes, and exist in a closed spiro (colorless) form featuring an absorbance at <400 nm, which undergoes a ring-opening rearrangement upon UV exposure to an open merocyanine (colored) form with an absorbance from 500 to 700 nm (Figure 6.9a) [151]. Photoswitchable spironaphthoxazine (NISO) has been conjugated to tetramethylrhodamine (TMR) in a FRET-based strategy to enhance intracellular imaging [159]. When NISO is present in its merocyanine (colored) form, it acts as a FRET acceptor for TMR diminishing its emission. However, when it is switched to its spiro (colorless) state, TMR donor emission increases due to diminished FRET. Spiropyran-based photochromic materials have been incorporated into a number of FRET formulations, including hyperbranched polymer micelles containing a hydrophobic fluorescent dye nitrobenzoxadiazolyl derivative [160], gadolinium-complexed materials for use in magnetic resonance imaging (MRI)-based deep tissue gene expression mapping [161], and QDs [162]. Using a spiropyran-based nitrospirobenzopyran (Nitro-BIPS)-conjugated fluorescent protein acceptor, FRET could be modulated with cycles of 365 and 546 nm light. This technique has been used to measure FRET efficiencies below 1% within a cell [157]. Also utilizing photochromic BIPS, a photoswitchable QD (psQD) has been developed for pcFRET [162]. By exposing the QD to white or UV light, the BIPS is transferred from colored merocyanine that acts as a FRET acceptor to colorless spiropyran that will not act as a FRET acceptor to the QD donor. In this way the emission of the QD can be modulated. Spiropyran-based dyes have also been incorporated into a number of luminescent NM-based probes for bioimaging [99,154,163,164]. Chen et al., for example, prepared FRET-based multicolor fluorescent and photoswitchable polymer NPs by incorporating two fluorescent dyes (EANI and NBDAA) (Figure 6.9b) and photoswitchable spiropyran into methyl methacrylate-based NPs via copolymerization (Figure 6.9b) [99]. By varying the ratios of the dyes, the emission signatures could be tuned such that the NPs exhibit multiple colors under a single excitation. Another increasingly common class of organic photochromic dyes includes the diarylethene-based photoswitchable compounds, which, like the spiro-based materials, undergo a structural open–closed photochromic transition [152,153,165–167]. j183 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3 Organic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.9 Spiro-based photochromic dyes. (a) The closed spiro form (SP – colorless) undergoes a ring-opening rearrangement upon UV exposure to the open merocyanine (MC – colored) form. (b) Generation of FRET-based multicolor photoswitchable fluorescent NPs by covalently incorporating two (EANI and SPMA) or three fluorescent dyes (EANI, NBDAA, and SPMA) under excitation at 385 nm [4-ethoxy-9allyl-1,8-naphthalimide (EANI) and allyl-(7-nitrobenzo[1,2,5]oxadiazol-4-yl)-amine (NBDAA)]. (c) Fluorescence emission spectra of three NP samples with different NBDAA feed (for samples NP-N1, NP-N3, and NP-N5, the NBDAA feed increased at a certain value) after visible light irradiation and UV irradiation. (d) Photograph of three NP dispersions (NP-N1, NP-N3, and NP-N5) after visible light irradiation and UV irradiation in the dark environment. (Reprinted with permission from Ref. [99]. Copyright 2012, American Chemical Society.) Diarylethene materials have been used to create photoswitchable dendrimers, where the diarylethene acts as a FRET acceptor that quenches the attached Cy3 donor emission via FRET, when switched off in the closed form [167]. Photochromic FRET imaging using this dendrimer was demonstrated within both HeLa cells and zebra Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 184 fish, with the potential for use in detailed imaging of cellular processes within cells or organisms. Photoswitchable QDs were developed by Diaz et al. by coating QD donors with an amphiphilic polymer containing diheteroarylethene acceptors and a spectrally separate Alexa647 dye to act as an internal standard by facilitating ratiometric measurements (Figure 6.10) [165,166]. The properties of this NP could make it potentially useful for intracellular imaging applications. Figure 6.10 Diarylethene-based photochromic dyes, and generation of photoswitchable QDs (psQDs). (a) The open diarylethene form (oPC – colorless) undergoes a ring-closing rearrangement upon UV exposure to the closed diarylethene form (cPC – colored). (b) Spectral signatures of the dual-color psQD components. Superposition of absorbance (solid lines) and emission (filled areas) spectra of PC, QD, and Alexa647, demonstrating the PC spectral overlap with the QD but not with the Alexa647. The spectra are normalized by their peak values. (c) Schematic of the dual-color psQD. The fluorescence of the QD is modulated by the photoconversion of the PC, while the Alexa647 fluorescence is constant. The PC in the open form (oPC) is photoconverted with UV irradiation to the closed form (cPC), which can then be photoreversed by direct excitation with visible light, or via FRET from the QD acting as a donor. (Reprinted with permission from Ref. [165]. Copyright 2012, American Chemical Society.) j185 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3 Organic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations In addition to organic dyes for pcFRET, photoswitchable FPs that can be encoded in the genome have also been demonstrated, and are discussed in more detail in Section 6.4.3 [155,168,169]. The use of pcFRET shows great promise for sensitive, high-resolution (diffraction unlimited) imaging of cellular processes, including rare protein interactions and individual protein movement within a cell. As a greater toolbox of photoswitchable materials for pcFRET continues to develop and researchers work on improving their inherent properties, pcFRET applications are sure to continue to grow. 6.3.7 Carbon Nanomaterials Carbon NMs represent a diverse class of materials with a variety of differing physical and chemical properties. This diversity is a direct result of a range of allotropic carbon material forms, including diamond, fullerene spheres and nanotubes, graphite and graphene, along with amorphous carbon. Of these many types, graphene-based sheets, nanodiamonds (NDs), luminescent carbon nanodots (Cdots), carbon NPs (CNPs), and carbon nanotubes (CNTs) possess relevant optical properties of interest for FRET applications. Of the graphene family of materials, graphene oxide (GO) sheets are reportedly fluorescence “superquenchers” that possess long-range energy transfer properties that make them ideal in FRET studies [170]. Most of the GO-based FRET sensors to date use DNA-based molecular recognition, in the form of molecular beacons, aptamers, or DNAzymes, for the specific detection of a range of target analytes from small molecules, such as heavy metals and mycotoxins, to proteins including thrombin, DNA, and even whole cancer cells [170–175]. Donor species range from traditional organic fluorophores to QDs and upconverting NPs (UCNPs); the diversity reflects the superquenching abilities of the GO materials. Wu et al., for example, demonstrated the simultaneous detection of the mycotoxins ochratoxin A (OTA) and fumonisin B (FB1) using two types of UCNPs modified with specific aptamers [173]. GO was used as the universal acceptor in the sensing scheme (Figure 6.11). In the absence of the target mycotoxins, the aptamer-modifed UCNPs interacted with the GO surface resulting in FRET and effective quenching of the UCNP luminescence. Addition of the mycotoxins, which bind to the aptamermodifed UCNP, altered the GO–aptamer-modifed UCNP interaction, resulting in an off–on sensor whose resulting luminescence spectra was mycotoxin specific. Dependent on the synthetic approach, GO sheets, especially the reduced form, in suspension or solid thin films can exhibit luminescent properties and have been used as donors in combination with Au NP acceptors for FRET detection of DNA hybridization and microcystins [176–178]. Graphene sheets smaller than 10 nm have also been found to possess photoluminescent (PL) properties [including upconversion (UC) and downconversion PL], and are referred to as graphene quantum dots (GQDs) [179,180]. Fan et al. found that 2,4,6-trinitrotoluene (TNT) effectively quenched GQDs luminescence via FRET upon the p–p stacking interaction that occurs between the two species [181]. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 186 Figure 6.11 GO sheets as universal quenchers in FRET-based assays. (a) Preparation of aptamers–UCNPs for mycotoxin detection. (b) Schematic illustration of the multiplexed upconversion FRET bioassay using aptamers–UCNPs (donors) and GO (universal acceptor) for FB1 and OTA detection. (Reprinted with permission from Ref. [173]. Copyright 2012, American Chemical Society.) CNTs comprise graphene tubes and, like the sheets, can either possess luminescent or superquenching properties depending upon morphology, synthesis, and purity. In general single-walled CNTs (SWCNTs) and sometimes double-walled CNTs (DWCNTs) are found to have luminescent properties [182], while multiwalled CNTs (MWCNTs) are considered superquenchers [183]. CNTs have successfully been used as donors and acceptors in FRET applications where they have been coupled with traditional organic dyes, QDs, and even lanthanide ions [184–187]. For example, QD-labeled ssDNA is found to undergo a strong interaction with CNTs, resulting in significant quenching of the QD luminescence [187]. Binding of the target influenza A virus DNA resulted in a significant decrease in the DNA–CNT interaction and an increase in QD emission, with a limit of detection (LOD) of 9.4 nM and excellent single-base mismatch discrimination. Synthesis and purity of the CNTs appear to be a key requirement for their successful application in FRET. While studying the fluorescence quenching of the dyes dansyl hydrazine and panacyl bromide covalently attached to SWCNTs, Chiu et al. found that the panacyl bromide quenching, unlike dansyl hydrazine, was very sensitive to the CNT purification method, specifically the metal impurities left over from CNT manufacture, suggesting care should be taken when interpreting data [185]. As an alternative to CNTs and GO sheets, which can be quite large as discrete labels, CNPs, NDs, and C-dots are relatively small and compact labels that are j187 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.3 Organic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations starting to find application in FRET [188–191]. CNP acceptors (quenchers) coupled with UCNPs have been applied to the detection of thrombin, using aptamers, and matrix metalloproteinase-2 (MMP-2), using a polypeptide substrate [192,193]. NDs, whose intense fluorescence properties arising from nitrogen-vacancy (NV) point defects in their nanocrystalline structure, have been investigated as donors in FRET studies with a number of near-IR dyes [194–196]. C-dots that are sub-10 nm particles that become fluorescent upon surface passivation have yet to be applied to FRET applications, but like NDs have great potential as donors. Many of these carbon NMs are still relatively new and studying their inherent physical properties and understanding the mechanisms that govern them is still very much a work in progress. This is hampered somewhat by the fairly complex and normally poorly controlled methods typically used to generate carbon NMs, such as chemical vapor deposition, electric arc discharge, or laser ablation. These methods typically produce a range of products containing a variety of impurities that have to undergo some type of purification in order to obtain the desired end product. That said, there are an increasing number of manufacturers who offer various carbon NMs (e.g., Sigma-Aldrich, Carbon Solutions Inc., Nano-C1, Microdiamant, and NanoAmor: Nanostructured and Amorphous Materials Inc.), oftentimes premodified with functional groups that aid in solubility and bioconjugation, which may encourage more widespread application. 6.4 Biological Materials Biological materials and biologically inspired materials (i.e., nonnatural amino acids), similar to the organic materials, are a diverse group including, molecules, proteins, and protein complexes. In addition, biological reactions creating bio- or chemiluminescence are an interesting alternative to the requirement for an external excitation source. 6.4.1 Natural Fluorophores Of the various naturally occurring fluorophores, including certain amino acid residues, reduced nicotinamide cofactors (NADH and NADPH), flavins (FAD and FMN), porphyrins, and pyridoxal derivatives, it is the aromatic amino acids, Trp, tyrosine (Tyr), and Phe shown in Figure 6.12 that dominate FRET applications [47]. The primary advantage to using these naturally occurring amino acids is that they have an endogenous presence in proteins, and when combined with FRET analysis can be used to study protein structure and dynamics [197]. Imaging studies of proteins can thus be completed with limited to no modification and even if these residues are not present in a protein, they generally can be incorporated into the peptide sequence with a minimal effect on its size, structure, and subsequent interactions. The strong UV absorbance of proteins at 280 nm (commonly used for Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 188 Figure 6.12 Naturally fluorescent amino acid residues. Structures of the aromatic amino acids, tryptophan (Trp), tyrosine (Tyr), and phenylalanine (Phe). quantitation) as well as an emission at 340–360 nm, originate mostly from the indole ring of Trp; Tyr and Phe contribute to a much lesser extent [47]. The negligible QY (0.02) of Phe makes it less amenable to FRET, except perhaps in intraprotein configurations. Tyr is prone to quenching and energy transfer to Trp, leaving Trp as the most reliable residue for FRET (detailed in Refs [47,50,198]). A potential liability in the use of Trp for FRET is that the excitation lines and any donor/acceptor dyes will be confined to the UV region. The fluorescence from these residues is also environmentally sensitive and so their placement deep within a protein structure will produce results that differ from those at the terminus of a small peptide. The primary application of endogenous fluorophores, in particular Trp, is to study the structure and function of proteins and peptides. FRET exchanges often occur between Trp and other fluorophores [such as nonnatural amino acids (see Section 6.4.2), or organic dyes], but FRET can also occur between two Trp residues, termed homo-FRET. Because of the low QY of Trp and its random distribution in molecules of interest, this homo-FRET rarely occurs. However, Kayser et al. found it to be a particularly useful tool for studying structure–function relationships in monoclonal antibodies, which were found to have an unusually high Trp content [199]. Another use of Trp as a natural fluorophore for FRET is to probe protein folding/unfolding, an important yet poorly understood biological process [200–202]. For example, Jha et al. used FRET to study the unfolding of a small protein, monellin [202]. The transition of this small protein, from an unfolded to a folded state, is not completely understood but by using a naturally occurring Trp and additionally labeling the protein with the FRET acceptor thionitrobenzoate, researchers were able to determine that monellin unfolds gradually rather than all in one motion [202]. Visser et al. utilized Trp homo-FRET to characterize protein folding of apoflavodoxin, demonstrating the potential of this method as a powerful means to understand some of the still unknown mechanisms of protein folding [201]. Another salient example leveraging Trp’s fluorescent properties to determine protein structure and function is its use in understanding the membrane transport protein LacY [203]. LacY is a sugar/Hþ symporter of E. coli bacteria and is widely studied. Trp fluorescence or quenching was used to determine the conformation of LacY and the function of the j189 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.4 Biological Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations symporter. Understanding protein structural variations could also aid in determining the pathogenic process of certain diseases. Lee et al., studying the human a-synuclein protein associated with Parkinson’s disease, used FRET between a fixed Trp donor and a modified 3-nitrotyrosine acceptor to demonstrate an elongated structure for the mutant protein associated with the disease [204]. While amino acids tend to dominate endogenous fluorophore FRET research, porphyrins, which physiologically form transition metal complexes [e.g., iron (see Section 6.5.2)], are found to have strong luminescent properties and have been demonstrated as FRET acceptors and donors in various formats. Lovell et al., for example, employed a caspase-3-specific peptide sequence modified with a rhodamine donor and porphyrin acceptor to monitor caspase activation in single cells following induction of cell death [205]. In the pursuit of artificial light-harvesting systems, porphyrin donors immobilized onto a clay surface (acceptor) were found to reach energy transfer efficiencies approaching the ideal 100% [206]. 6.4.2 Nonnatural Amino Acids Intrinsic natural probes such as Trp, Tyr, and Phe are highly useful for visualization of protein structure, movements, and interactions, and as mentioned are either naturally occurring or can be introduced without significantly impacting the protein structure. However, due to low QYs of both Phe and Tyr, Trp is the only widely used natural fluorescent probe. To overcome this limitation nonnatural (also called unnatural or noncanonical) amino acids have been fabricated [207–213]. These nonnatural amino acids can provide larger QYs and new FRET pairs for protein structure function analysis. One recently developed fluorescent nonnatural amino acid p-cyanophenylalanine (PheCN) can be incorporated into a protein, minimally disturbing the native protein structure [214], and can act as a FRET donor to Trp (Figure 6.13) [207]. The photophysics of this useful nonnatural amino acid has recently been characterized [211,215]. The Trp–PheCN FRET pair has been used to determine detailed protein folding and unfolding in two small proteins, the villin headpiece subdomain (HP35) and the lysin motif (LysM) domain [200]. Two other nonnatural amino acids, 7-azatryptophan (7AW) and 5-hydroxytryptophan (5HW) (Figure 6.13), were recently found to be FRET acceptors to PheCN [210]. The 7AW–PheCN FRET pair had a greater separation of fluorescent spectrums than PheCN–Trp. Moreover PheCN, Trp, and 7AW can be used in a multistep FRET system to investigate interactions of three points on a protein (see Section 6.6) [210]. Even newer nonnatural FRET pairs are continuously being developed. Recently, L-4-cyanophenylalanine (pCNPhe) and 4-ethynylphenylalanine (pENPhe) were used as FRET donors to Trp in order to probe the hydrophobic core of the protein T4 lysozyme [216]. Another nonnatural amino acid was created by combining p-aminophenylalanine derivatives with BODIPY fluorophores, generating a material with an emission wavelength greater than 500 nm [208]. These amino acids were incorporated into a calmodulin-binding peptide and FRET used to probe protein binding and resulting conformational changes. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 190 Figure 6.13 Nonnatural amino acids. (a) Structures of some selected nonnatural amino acids, p-cyanophenylalanine (PheCN), 7azatryptophan (7AW), and 5-hydroxytryptophan (5HW), derived from the naturally fluorescent amino acids with the chemical modifications shown in red. (b) Improving Stokes shift in eCFP using FRET with a nonnatural amino acid. Three-dimensional model of eCFP, which carries the fluorescent amino acid (1) at the surface-exposed position 39, is based on the crystal structure of eCFP (PDB entry 2WSN). Chemical structures of (bottom) the eCFP fluorophore 4-[(1H-indol-3-yl)methylidene] imidazolin-5-one (5; lex ¼ 434 nm, lem ¼ 476 nm) and the fluorescent nonnatural amino acid L-(7-hydroxycoumarin- 4-yl) ethylglycine (1; lex ¼ 360 nm, lem ¼ 450 nm) that together form a FRET pair. Normalized absorption and fluorescence spectra of 1 [Abs (1) and Flu(1)] and eCFP [Abs(eCFP) and Flu (eCFP)]. The major absorption band of eCFP shows considerable overlap with the fluorescence spectrum of 1, thus fulfilling a prerequisite for efficient FRET. (Reprinted with permission from Ref. [217]. Copyright 2011, American Chemical Society.) j191 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.4 Biological Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Nonnatural amino acids can also be used to modify FPs to enhance their fluorescent properties via intramolecular FRET [209]. The nonnatural amino acid L-(7-hydroxycoumarin-4-yl)ethylglycine, for example, was incorporated into recombinant cyan FP (CFP) [217]. The modified FP underwent FRET between the nonnatural amino acid (donor) and the FP’s natural chromophore (acceptor), resulting in emission at 476 nm with a 365 nm excitation wavelength (Figure 6.13). This large apparent Stokes shift of 110 nm is much greater than the natural 40 nm Stokes shift of CFP alone. Nonnatural amino acids have also been applied to time-resolved FRET (TR-FRET), which can be used to evaluate protein folding dynamics. Historically, Trp has been used as a donor for TR-FRET measurements; however, a significant drawback is that it exhibits a high degree of variation in its fluorescent lifetime, depending on protein conformations. In an attempt to overcome this, an analogue of Trp, 5-fluorotryptophan (5F-Trp) has been proposed as a better candidate for TR-FRET [212]. The 5FTrp has more homogenous decay kinetics than Trp and is less environmentally sensitive, making it an ideal donor for TR-FRET for the determination of molecular structure in proteins. As new nonnatural fluorophores are designed, and the study of these and natural fluorophores progress, these materials could provide an even greater utility for understanding proteins at a molecular level utilizing FRET techniques. 6.4.3 Green Fluorescent Protein and Derivatives FPs represent an increasingly diverse class of fluorophores that have shown great potential as genetically encoded fluorescent tags for assessing protein location and function (monitoring protein–protein interactions) in cell studies and the development of in vivo (signaling dynamics such as calcium ions) and in vitro biosensors [4,22,168,218–221]. While GFP derived from the jellyfish Aequorea victoria represents the prototypical fluorophore of this protein family, various GFP mutations and Anthozoa (coral) homologues provide an increasingly diverse range of photophysical properties (Table 6.1 and Figure 6.14), which stem from their internal chromophores [155,168,169,218,222–224]. Newman et al. give an excellent monograph of the FP basics [168]. To briefly summarize, FPs self-generate their intrinsic chromophore from key internal amino acid residues, which nestle deep in the core of the characteristic 11-stranded b-barrel FP structure. The final photophysical properties of the mature FP are governed by the extent of p-conjugation, subsequent chromophore transitions, and interactions with the surrounding amino acid microenvironment with the chromophore (Figure 6.14) [155,168,169]. Key to the success of FPs has been the ability to genetically encode them via commercial plasmids that can be expressed in a variety of cells/organisms (available from ClonTech Laboratories, Inc., Life Technologies, Evrogen, and MBL Intl. Corp.). The genetically encoded FPs are commonly attached to Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 192 Table 6.1 Properties of some representative FPs summarized from www.microscopyu.com and Ref. [168]. Protein Source In vivo structure A. victoria E. quadricolor D. striata A. victoria A. victoria E. quadricolor Monomera) Monomer Monomer Monomera) Monomera) Monomer A. victoria Copepod sp. A. victoria E. quadricolor Clavularia coral Emission max. nm Extinction coefficient M – 1 cm – 1 Quantum yield 383 399 402 439 435 458 445 456 467 476 477 480 29 900 52 000 51 000 32 500 35 000 37 000 0.31 0.63 0.48 0.4 0.51 0.57 Monomera) Dimer Monomera) Monomer Monomer 395/475 482 484 483 493 509 502 507 506 509 21 000 70 000 56 000 56 500 70 000 0.77 0.53 0.6 0.6 0.8 E. quadricolor A. victoria A. victoria A. victoria Monomer Monomera) Monomer Monomera) 508 514 516 517 524 527 529 530 64 000 83 400 77 000 104 000 0.6 0.61 0.76 0.77 D. striata D. striata Cerianthus sp. E. quadricolor Monomer Monomer Tetramer Dimer 540 548 548 553 553 562 573 574 6000 71 000 60 000 92 000 0.7 0.69 0.64 0.67 D. striata E. quadricolor D. striata A. sulcata D. striata D. striata Tetramer Monomer Monomer Tetramer Monomer Monomer 558 555 568 576 574 587 583 584 585 592 596 610 75 000 100 000 38 000 56 200 90 000 72 000 0.79 0.48 0.3 0.05 0.29 0.22 D. striata D. striata E. quadricolor E. quadricolor Monomer Monomer Monomer Dimer 598 590 600 605 625 649 650 670 86 000 41 000 67 000 70 000 0.15 0.1 0.2 0.06 Absorbance max. nm Blue EBFP TagBFP mBlueberry2 ECFP CyPet TagCFP Green GFP (wt) Turbo GFP EGFP TagGFP2 mWasabi Yellow TagYFP EYFP mCritrine Ypet Orange mBanana mOrange OFP TurboRFP Red DsRed TagRFP mTangerine AsRed2 mStrawberry mCherry Near-IR mRaspberry mPlum mNeptune eqFP670 a) Weak Dimer formation. b) E ¼ Entacmaea genus. c) D ¼ Discosoma genus. j193 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.4 Biological Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.14 Fluorescent proteins. (a) Structure of A. victoria GFP showing the dimensions of the protein, the intrinsically derived p-HBI chromophore, and several key residues surrounding the chromophore (image generated using PyMOL open access and PDB ID 1w7s). (b) Chromophore structures of representative FP color variants within each spectral class. The conjugated ring structure of each chromophore is colored according to its emission profile. (Reprinted with permission from Ref. [168]. Copyright 2011, American Chemical Society.) proteins of interest through the creation of FP chimeras that can subsequently be used in cellular studies of protein location, protein–protein interactions, and the development of biosensors to monitor cell signaling processes [4,168,219,220]. In addition to the benefit of genetic manipulation, there are a number of advantages and disadvantages to the use of FPs as fluorescent tags that should be factored into their use as FRET donors/acceptors. FPs have a wide range of QYs (see Table 6.1), ranging from 0.04 for AQ143 to 0.91 for ZsGreen, which depend on Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 194 the mutations present and final chromophore structure, although the majority are generally good with QYs > 0.5 [168]. Certain FPs have been found to possess two-photon absorption properties that can be very advantageous in deep tissue imaging applications [225,226]. Photophysically, FPs can take several hours to fully mature as the final chromophore is formed through the protein folding process and any subsequent chemical transitions [155,169,168,221]. The relative brightness of their fluorescence intensity is found to be intimately linked to the efficiency of the FP folding process and mutation time [155,168,221]. FPs generally have relatively broad absorption and emission profiles that may preclude “multiplex” analysis. FPs are also prone to photobleaching and have known susceptibilities to pH, temperature, O2 concentration, and other environmental conditions. FPs are of a fairly large size from a fluorescent tag point of view, Mw 25–30 kD and upwards, which can be problematic in terms of maintaining the desired function of the labeled target protein. In addition, certain FPs have a tendency toward the formation of oligomers (dimers and tetramers), which can further confound the size issue [168]. Also, location of the FP-tag in the target protein must be carefully chosen so as not to significantly impact FP maturation and therefore brightness [168]. Researchers continue to develop FP mutations that attempt to address a number of these issues, including improved photostability [223], decreased oligomerization [222], emission in the near–far IR region [169,227], improved Stokes shifts [155], or photoactivatable (including photochromic) properties [155,168,169,228]. Use of FP pairs in FRET and bioluminescence resonance energy transfer (BRET) (see Section 6.4.6) applications is ever expanding, with dramatic implications for in vivo imaging, biosensors, and cellular studies in particular [22,155,168,219,229–235]. FPs have revolutionized the detection and study of cellular events, and the more recent combination of FPs and FRET imaging has taken this detection to new levels of precision, allowing the study of protein–protein interactions and tracking biochemical and protein signaling dynamics, reviewed in a number of excellent publications [22,155,168,219,220,229,232,236]. There are two main FRET–FP strategies employed when studying cellular processes, the nature of which is dependent on the process being investigated. In the first strategy, typically employed for measuring intermolecular protein–protein interactions, it is common to tag each protein with either the donor or acceptor FP (Figure 6.15a), as FRET emission will only take place when the two (or more) proteins of interest interact. In the second strategy a protein or biosensor construct is labeled with both the donor and acceptor FPs, and interaction with the target species of interest results in some measurable change in the FRET signal. There are a number of different biological processes that can be monitored using the second strategy, including protease activity and activation, Ca2þ ion fluctuations, measuring second messengers such as cyclic adenosine monophosphate (cAMP) and cyclic guanosine monophosphate (cGMP), studying phosphoinositide dynamics, and G protein-coupled receptor (GPCR) activation, and hence a wider range of formats are/can be employed, some of which are highlighted in Figure 6.15b, and recently reviewed in Refs [22,168,236]. j195 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.4 Biological Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.15 Representative FRET-based sensor formats that incorporate FPs. (a) Monitoring protein–protein interactions. Here, each protein is labeled with a FP that upon interaction results in FRET. (b) Biosensors in which binding of a small molecule induces the association of two distinct moieties within the single polypeptide chain. (c) Biosensors for posttranslational modification. (d) Biosensors in which a protein undergoes a conformational change upon binding its small-molecule ligand. (e) Biosensors for protease activity. The donor is CFP and the acceptor is YFP in these representations, however, a variety of other FP FRET combinations could be substituted. (Reprinted with permission from Ref. [22]. Copyright 2009, American Chemical Society.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 196 While the use of FP pairs in FRET has been constantly expanding, advances need to be made continually to enhance its effectiveness. Shortcomings in FP–FRET include a large range in brightness in FPs, cross talk due to large emission spectra of FPs, slow development of intramolecular sensors, and limited ability to create stably transfected cells lines with FRET FPs. Though recent advances have achieved some success in expressing FP–FRET pairs in a cell line [237,238], difficulty still exists in achieving this goal [236]. With advances in developing cell lines to express FP–FRET pairs, more research can be conducted using these systems, as researchers will be able to more readily conduct experiments without the difficulty of transiently transfecting cells. Brightness and cross talk issues can be mitigated by further improving FPs for FRET. Development of additional FPs for FRET has also led to multiparameter imaging using dual FRETpairs. Development of this technology has important research implications as entire signaling cascades could potentially be imaged in the same cell. The most commonly used FRET pair is CFP–yellow FP (YFP), but even with this pair significant cross talk exists [229]. New, recently developed FPs include enhanced GFP (EGFP) and mCherry, which have similar brightness and reduced cross talk compared to CFP–YFP [239]. Lam et al. demonstrated a Clover–mRuby2 FP–FRET pair combination for monitoring kinase acitivity, guanosine triphosphate hydrolase (GTPase) activity, and transmembrane voltage with significantly improved photostability, FRET dynamic range, and emission ratio changes versus CFP–YFP [224]. An orange fluorescent protein with a large Stokes shift (LSSmOrange) has been developed, which can be used for intracellular imaging, potentially allowing two FRET pairs in combination with CFP–YFP [240]. In a salient example of multiparameter imaging, four different cellular events were recorded simultaneously using FP–FRET imaging [241]. Here, two CFP–YFP FRET sensors that could be spatially resolved were combined with a spectrally distinct mCherry–mORange FRET pair and a fourth sensor Fura Red. The development of new FRET protein pairs for multiple parameter imaging could greatly expand the use of FP–FRET and will accelerate discovery of cellular processes [242], cancer research [236], and toxin detection [243]. Photoswitchable FPs, a subset of the larger FP community, for pcFRET have also been demonstrated and have great potential for high-resolution imaging applications [155,168,169]. Photochromic/photoswitchable FPs are under continuous development, and the various mechanisms/factors that govern photoswitching have recently been reviewed [155,168,244]. Switching is thought to occur primarily through a cis–trans isomerization of the FP chromophore, as demonstrated with isolated synthetic chromophore analogues [245,246], however the chromophore environment within the FP structure also plays a key role in determining a FPs switchability [78]. Recently, it has been discovered that substitution of certain key amino acids in the chromophore environment within the FP structure can improve and/or restore the photochromic behavior of the FP chromophore, leading to improved photoswitchable FP mutations for pcFRET studies [78,244,247]. Grotjohann et al., for example, generated a reversibly switchable EGFP (rsEGFP) mutant that could be cycled 1200 times before experiencing a 50% reduction in j197 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.4 Biological Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations fluorescence, in stark contrast to FP Dronpa that experiences a 50% reduction after only 10 cycles [247]. The red FP (rsTagRFP) can be photoswitched from “on” and “off” fluorescent states using 445 and 570 nm lights, respectively [248]. The rsTagRFP was subsequently used as an acceptor with an enhanced YFP (EYFP) donor to monitor protein interactions in pcFRET studies of living cells, where the on–off switching provided confirmation of the FRET signal and the protein interaction. Other photoswitchable FPs include Dronpa [249] (and variants bsDronpa and Padron [228]), rsCherry [250], and rsCherryRev [250]. While pairing two FPs for FRET is the most common combination when utilizing FPs, there are increasing examples of FP donors or acceptors being paired with other fluorescent materials for FRET applications. There are a number of BRET examples discussed in Section 6.4.6 and also multi-FRET examples where the FP is coupled with light-harvesting complexes or proteins (see Section 6.6). Rice created a kinesin C-terminal GFP fusion and labeled the kinesin with tetramethylrhodamine, allowing FRET monitoring of protein conformational changes upon binding nucleotides [251]. Hoffman dual labeled a GPCR system with CFP and the FlAsH system to monitor receptor activation demonstrating that, unlike the equivalent CFP–YFP-labeled system, downstream signaling was not disrupted by the FlAsH acceptor [252]. FP acceptors combined with QD donors are becoming an increasingly common combination, especially in the development of biosensors for measuring protease activity [230] and intracellular pH (Figure 6.16) [79]. Clearly, FP-based FRET has already made a significant impact on our understanding of cellular processes, and as the materials themselves continue to evolve and improve, more sophisticated applications can be expected. I Figure 6.16 QD–FP FRET-based pH sensor. (a) Schematic demonstration of the pHdependent energy transfer between the QD and the FP. In an acidic environment, energy transfer to the FP FRET acceptor is minimal, yielding a high QD signal; at neutral or basic pH, energy transfer is more efficient, producing an enhanced FRET signal. (b) A pH titration of QD–FP probes containing the FP acceptor mOrange M163 K showing increased energy transfer at alkaline pHs with a clear isosbestic point. Cellular imaging of QD–mOrange pH sensor. (c) Schematic of probe color changes during progression through the endocytic pathway. FRET efficiency is high in the neutral pH of the extracellular environment and early endosome. FRET efficiency decreases as the endosome matures and the endosomal pH drops, resulting in diminished emission from mOrange and recovery of some QD signal. Any probe that escapes the endosome regains its elevated FRET efficiency in the pH neutral cytoplasm. (d) Fluorescence microscopy images immediately after delivery of the probe and 2 h post delivery. The QD images (left) demonstrate consolidation of the probe in the endosomes over time; images of the direct excitation of mOrange (center) and FRET emission (right) indicate a clear decrease in the mOrange emission and the FRET efficiency of the probe with maturation of the endosome. (Reprinted with permission from Ref. [79]. Copyright 2012, American Chemical Society.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 198 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j199 6.4 Biological Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 6.4.4 Light-Harvesting Proteins Phycobiliproteins (PBPs) are the colored component proteins found in naturally occurring phycobilisome-based light-harvesting complexes (see Section 6.6) [253,254]. Their intense color originates from the combination of chromophores they contain, termed phycobilins. Examples of phycobilins include phycocyanobilin (PCB – blue), phycoerythrobilin (PEB – red), phycourobilin (PUB – yellow), and phycobiliviolin (PXB – purple). In phycobilisome-based systems, there are four main PBPs: allophycocyanin (APC), phycocyanin (PC), phycoerythrin (PE), and phycoerythrocyanin (PEC); some spectroscopic details are summarized in Table 6.2 [254,255]. Peridinin–chlorophyll–protein (PerCP or PCP) is a lightharvesting complex found in dinoflagellates and has a smaller MW (35 kDa) than most of the 100–240 kDa PBPs mentioned, which can be beneficial when the size of the fluorescent label is a concern. Since their inception as fluorescent probes, the purified PBPs themselves (e.g., phycofluor probes), complexes containing multiple PBPs (e.g., PBXL), or the smaller PerCP materials have become common fluorescent labels in bioassays, especially in flow cytometry [254,256–259]. Their use as fluorophores in FRET has been more limited to date, and may be due in part to their relatively large size – although some of the smaller materials being developed (e.g., PerCP and CryptoFluorTM) may address this concern. That said, they have been demonstrated in sandwich-based peptide and antibody FRET assays for target analyte detection and for demonstrating FRET in combination with Au and QD NMs [121,257,260–264]. The combination of APC (acceptor) and Eu-cryptate (donor) has been used for the time-resolved FRET-based detection of telomerase activity [262], inhibitors of hepatitis C virus core dimerization [264], and small-molecule inhibitors of HIV-1 fusion [263]. A number of companies sell PBP- and PerCP-based materials, including Life Table 6.2 Properties of some representative phycobiliproteins summarized from www.columbiabiosciences.com and Ref. [254]. Protein Allophycocyanin (APC) R-Phycoerythrin (RPE) B-Phycoerythrin (RPE) Approx. Types of Approximate MW phycobilins number of kDa present phycobilins 100 240 240 PCB PEB and PUB PEB and PUB 6 34 34 Extinction coefficient M – 1 cm – 1 Quantum yield 657.5 2.4 105 0.68 573 6 1.96 10 0.84 572 2.41 106 0.98 Absorbance Emission max. nm max. nm 652 625a) 565 498a) 545 563.5a) Note: Exact properties are dependent on the origin of the protein PCB: phyocyanobilin, PEB: phycoerythrobilin, and PUB: phycourobilin. a) Additional absorbance peak. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 200 Technologies, Jackson ImmunoResearch (mainly streptavidin antibody-labeled and secondary antibody-labeled materials), the SureLight1 series from Columbia Biosciences, and the CryptoFluor materials from Martek Biosciences Corp. (sold by Sigma-Aldrich). 6.4.5 DNA-Based Macrostructures/Nanotechnology DNA is an incredibly complex polymeric material comprising four monomers – adenine (A), thymine (T), guanine (G), and cytosine (C) – that can be single stranded (ss) or doubled stranded (ds), which results when two complementary strands hybridize [265]. Researchers are increasingly interested in using the polymeric nature of DNA to synthesize unique 2D and 3D DNA-based macromolecular/nano structures [265–268]. The use of functionalized DNA nanostructures for light harvesting and charge separation has recently been reviewed [269]. Researchers have already demonstrated the potential of these DNA structures as fluorescent labels using some relatively simple DNA constructs. Accumulation of perylene- and pyrene-based fluorophores in DNA duplexes has been shown and produces fluorescent excimer structures very strongly, which possess large Stokes shifts, making them ideal donors in FRET applications [270,271]. Kumar and Duff engineered some unique DNA–protein complexes that demonstrated potential as light-harvesting complexes [272]. The Armitage group developed DNA tetrahedron and duplex fluorescent nanotags using FRET to shift the emission of the DNA nanotag further into the red region relative to the donor dye alone (Figure 6.17) [273–275]. Although the majority of these DNA-based structures are designed in-house, the DNA sequences themselves are often synthesized and purchased commercially [ from companies such as Integrated DNA Technologies (IDT)]. Genisphere1 sells a 3DNA-based dendrimer that is marketed for signal amplification in a number of bioassays. Given the ease with which fluorescent dyes can be incorporated into these unique structures, through either intercalation or covalent attachment, it seems likely that the utility of these materials as fluorescent labels and their subsequent use in FRET-based applications will increase in the future. 6.4.6 Enzyme-Generated Bioluminescence Enzyme-generated bioluminescence (BL) is used in a particularly advantageous variant of FRET known as BRET. BL is a naturally occurring phenomenon found in certain beetles and bacterial or marine species, where various substrates (luciferins) react with enzymes (luciferases) in the presence of O2 (and sometimes other cofactors) to produce light emission (Table 6.3) [276–281]. The exact wavelength of the light emission is found to be dependent on a number of factors, j201 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.4 Biological Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.17 DNA NMs for FRET. (a) Assembly of DNA tetrahedron nanotags. Four strands with partially complementary sequences form the DNA tetrahedron nanostructure template for the self-assembly of intercalator dyes. Black sections represent two-nucleotide long, singlestranded hinges. (b) Schematic description of ET (energy transfer) in a tetrahedron nanotag loaded with YOYO-1 intercalated dyes and covalently attached Cy3 acceptor dyes. (c) Fluorescence emission of tetrahedron nanotags with 0–4 covalently attached Cy3 molecules. Spectra acquired by excitation at 440 nm. Samples contained 50 nM DNA tetrahedron and 1.28 M YOYO-1. ET efficiencies given in legend were determined by the percentage of decrease in the YOYO-1 emission at 509 nm. (Reprinted with permission from Ref. [274]. Copyright 2009, American Chemical Society.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 202 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License including the structure of the luciferin, the nature of the luciferase, and the presence of accessory proteins, physiologically, GFP or its derivatives [277]. BL itself has found application as a reporter in many bioassays, both in vitro and in vivo [278–282]. The use of the BL reaction as a donor within energy transfer mechanisms is, in fact, an intrinsic process observed in many sea creatures such as A. victoria (jellyfish) and Renilla reniformis (sea pansy), where accessory proteins (e.g., GFP) modify the color of the emission through BRET [277]. Researchers have used BRET in a number of applications, most notability for studying protein–protein interactions, leading to drug discovery, and increasingly for biosensing and in vivo imaging [283–288]. As with FRET, BRET depends upon spectral overlap between the donor emission and the acceptor absorption, and it is similarly efficient over distances up to 10 nm [278,284,287]. The principal advantage of BRET is removal of the excitation source, negating problems such as light scattering, high background noise, direct acceptor excitation, photodamage to cells, and photobleaching effects [287,289]. The most commonly exploited luciferases for BL are the eukaryotic firefly and Renilla luciferases (Rluc), with the wild-type enzymes generating blue-green emission. DNA vectors with the desired luciferase gene or plasmids can be purchased through a variety of sources such as Promega, New England Biolabs, and Targeting Systems. These vector complexes are then internalized by the cell of choice, where the luciferase gene can be transcribed by ribosomes to produce the desired enzyme. Luciferases can also be fused to other proteins or fragmented to monitor protein interactions of interest [276,290]. Improving the BL properties of these systems is an active field, and research ranges from generating a wider variety of BL colors, spanning the visible to nearinfrared wavelengths, to improving the emission kinetics of the BL, by increasing the intensity and/or decay of half-lives (reviewed in Ref. [278]). Research in this area is two pronged, focusing on the luciferases themselves and the luciferin substrates, with a wide range of protein mutants and substrate analogues now produced/utilized (reviewed in Ref. [278]). Sun et al. recently reviewed progress in D-luciferin amino analogues that produced emissions ranging from 460 to 609 nm with wild-type luciferase [291]. Other D-luciferin analogues, reviewed by the Meroni group, give insight into a variety of chemical manipulations that may result in altered emitted light [292]. Caged luciferin, for example, is an alternative firefly luciferase substrate designed for intracellular delivery available from Molecular Probes [293]. Caged luciferin has also recently been synthesized for the real-time in vivo imaging of H2O2 production in living mice [294]. Coelenterazine substrates have similarly been altered to provide both enhanced brightness and enhanced duration of photon emission, including ViviRenTM and EnduRenTM from Promega, and several analogues with different emissions are available from Molecular Probes and Biotium [295,296]. In an effort to continuously improve BRET, recent studies by Zhang and coworkers have shown improvements in both sensitivity and limit of detection by 10-fold and 7-fold, respectively, through the use of an enhanced buffer environment [297]. j205 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.4 Biological Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations As illustrated in Table 6.3, a variety of luciferase types can be combined with essentially three different chemical substrates to produce BL for BRET applications. Bacterial luciferases, for example, catalyze the oxidation of reduced flavin mononucleotide (FMNH2) and a long-chain aliphatic aldehyde in the presence of O2, yielding blue light [298,299]. Bacterial luciferase, however, is not optimized for most mammalian cell lines and the substrate, FMNH2, is rapidly oxidized in air, generating short bursts of light and not a steady BL emission [300]. Firefly luciferases differ from bacterial luciferases in that they require an additional cofactor, ATP, in order to catalyze the oxidation of the substrate luciferin, leading to the emission of green-yellow light. Initial BL in this case is intense and then decays to a low-sustained luminescence, which can be aided by an additional cofactor, coenzyme A, to yield more stable, high-intensity luminescence [301]. Rluc is one of the most commonly utilized luciferase systems and uses the substrate coelenterazine in the presence of O2. It has been widely adapted to mammalian cell lines, and hRluc is available from a variety of sources, including Promega and PerkinElmer Life Sciences; cofactors are not needed. Improvements have been made upon the wild-type enzyme, producing the often used mutants, namely Rluc2 and Rluc8, which have improved BL properties [287,302,303]. Another coelenterazine-based luciferase that has begun finding wide application is derived from Gaussia princeps (Gluc or hGluc) and marine copepods (BL crustacean), and has been optimized for expression in both bacterial and mammalian cells [304,305]. The low molecular mass of Gluc (20 kDa) compared to Rluc (36 kDa) addresses problems associated with steric constraints in chimeric protein fusions. Gaussia luciferase expressed in mammalian cells reportedly generates light up to 1000-fold brighter than that of native Renilla [306,307]. BL photoproteins from jellyfish and hydroid species, namely aequorin (from A. victoria) and obelin (from Obelia longissima), respectively, are also coelenterazine-based enzymes that differ from luciferases in that the enzyme is complexed to its coelenterazine substrate and is Ca2þ sensitive [308]. The principal application of these photoproteins has been as Ca2þ reporters [309,310]. In BRET applications aequorin has been employed to monitor the protein–protein interactions between SA and a biotin carboxyl carrier protein [311]. In terms of BRET these BL enzymes are most commonly coupled with FPs, a combination that has been fueled by the desire to push the BRET emission further into the near-IR for optimal in vivo imaging. This has been facilitated by the increasing number of mutant BL enzymes, substrate analogues, and mutant FPs that allow good spectral overlap of the generated BL with the FP absorption [155,218,278]. As a result there are a variety of BL enzyme–FP BRET configurations, termed BRET x [288,302,312,313]. In BRET1 the BL enzyme variant is Rluc or Rluc8 and the accepting protein is a GFP variant, YFP. Oxidation of the substrate coelenterazine-h by Rluc results in BL with a 480 nm peak, which, through energy transfer to YFP, generates a fluorescent emission peak at 530 nm. Typical uses of BRET1 are ligand screening applied to real-time detection of protein–protein and protein–ligand interactions, such as agonist-induced interactions of the GPCRs family of receptors [287,288,314]. BRET has been used to study a number of the Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 206 GPCRs, irrespective of their G protein-coupling selectivity [287,288], and is particularly amenable to high-throughput formats [287]. BRET2 utilizes the BL enzyme variants Rluc, Rluc2, or Rluc8 and acceptor GFP2 (a GFP variant with Ex 400 nm and Em 511 nm). Reaction of the coelenterazine analogue substrate, DeepBlueC (coelenterazine 400A, sold commercially by Biotium, Inc., PerkinElmer, and NanoLight Technology), with the Rluc enzyme causes a BL emission at 400 nm, resulting in excitation of GFP2 and an endpoint emission at 511 nm [287,288,314]. This configuration works like a standard BRET assay, but has a larger apparent Stokes shift resulting in more spectral resolution between the donor–acceptor pair. This technique was successfully applied to visualize protein–protein interactions in mice [312] and RNA detection and quantification (Figure 6.18a) [315]. BRET3 again uses Rluc, Rluc2, or Rluc8 and the fluorescent protein mOrange with a coelenterazine substrate [313]. Advantages of the BRET3 combination include several-fold improvement in light intensity, as well as improved spatial and temporal resolution for measuring intracellular events in a single cell. This improved BRET strategy allows the visualization of protein–protein interactions within small living animals [287,288,313,314]. The Gambhir group has subsequently demonstrated additional BRET x systems with various combinations of Rluc variants (Rluc8 and Rluc8.6) combined with two red FPs, TagRFP, and TurboFP635, using the substrates coelenterazine and its analogue coelenterazine-v to generate a red light-emitting 600–650 nm reporter system for in vivo protein–protein association studies [302]. BRETx systems are initially characterized/optimized by generating a fusion protein of the BL enzyme and the FP, and these fusion proteins represent interesting tags in their own right and may be useful labels for a variety of bioassays [316]. For example, they have been used in sequential BRET–FRET assays, termed SRET, for detection of heteromerization in plasma membranes and ratiometric protease assays [286,314,317,318]. Extended BRET (eBRET) typically uses the enzyme variant Rluc or Rluc8 with YFP (Ex 480 nm and Em 530 nm). A presubstrate, a protected form of coelenterazine (EnduRen from Promega), is metabolized by endogenous esterases into coelenterazine-h similar to that used in BRET1. This provides a steady supply of substrate for luciferase oxidation within cells for extended periods of time up to 24 h [285,287,314]. Although less common, BL enzymes have also been coupled with traditional organic dyes and increasingly NMs, especially QDs as acceptors [283,286,319–321]. Currently QD–BRET has been used for sensing protease activity, protein–protein interactions and in vivo imaging, as reviewed in Ref. [283]. Enhanced BRET between the firefly Photinus pyralis luciferase variant PpyGRTS quantum rods (QRs) as the energy acceptor has also been described recently, with BRET ratios dependent upon enzyme loading, rod aspect ratio, and donor–acceptor distances (Figure 6.18b) [320]. Protease activity sensing in a QD–BRET system relies on a peptide substrate linkage between the luciferase and the QD that is cleaved in the presence of the protease of interest, reducing the BRET-based QD emissions [321]. QD–BRET has also been used to study protein–protein and receptor–ligand interactions. For example, Rao and coworkers fused luciferase (Luc8) to HaloTag protein and functionalized QDs with HaloTag ligands to study the resulting BRET that occurred when the HaloTag j207 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.4 Biological Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.18 (a) General strategy for BRETbased RNA detection. Probe1 consists of a 20mer oligonucleotide conjugated at the 50 end to the thermostable bioluminescent protein (RL8). Probe2 consists of a 20-mer oligonucleotide conjugated at the 30 end to the fluorescent protein GFP2. Both the probes are complementary to different portions of the same mRNA target gene. Thus, the target mRNA serves as a scaffold upon which the probes can bind, bringing the proteins into proximity to one another. When RL8 oxidizes its substrate, the energy produced is nonradiatively transferred to GFP2, which then emits photons at a characteristic wavelength as its chromophore returns to the ground state. A dual probe assay testing sensitivity with mixed populations of in vitro-transcribed cRNA as targets. RL8 and GFP2 were combined with various amounts of Fluc cRNA, while keeping the level of total cRNA constant by supplementing with nontarget cRNA. Statistically significant BRET signal was seen for as little as 1 mg Fluc cRNA. Inset: raw image obtained in IVIS-200. Rows from top to bottom match columns left to right of figure. Left image: shows GFP2 filter; Right image: shows RL8 filter. (Reprinted with permission from Ref. [315]. Copyright 2008, American Chemical Society.) (b) BRET between QRs and firefly luciferase enzymes. (i) BRET efficiency plots for PpyGRTS donors and QR acceptors. The summary of the BR measured with respect to aspect ratio at L ¼ 5 and 10 for (ii) CdSe/CdS and (iii) CdSe/ CdS/ZnS QRs. (iv) Illustration of the microstructure of the particular QRs studied, including dot-in-dot, rod-in-rod, and dot-in-rod types. (Reprinted with permission from Ref. [320]. Copyright 2012, American Chemical Society.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 208 protein irreversibly bound the HaloTag ligands [286]. Interestingly, in vivo imaging with a QD–BRET system has the potential to overcome the natural tissue scattering and autofluorescence, which affects a large number of BRET acceptors with short wavelength emissions. QDs are available with emission in the red and near-IR ranges, which coupled with their broad absorption profiles makes them an excellent acceptor for BRET. This was first shown in a mouse model where superficial and deep tissues were imaged using QDs with a 655 nm emission, which were coupled to Luc8 [319]. As highlighted later in the QD section (Section 6.5.5), the relatively narrow emission profiles open up the possibility of multiplexed sensing of protein– protein interactions using a number of QD–BRET pairs [286,321]. The BRET platform has been used for sensing a variety of analytes, including Ca2þ, through the use of the described photoproteins and ATP. BRET is possibly most often used for monitoring protein–protein interactions such as the earlier described GPCR activation as well as the QD–BRET system for determining reaction kinetics of HaloTag protein binding to HaloTag ligands [286]. Similarly, BRET has been utilized for nucleic acid hybridization assays and immunoassays, with the latter available for purchase from a variety of companies. As an example of a nucleic acid hybridization assay, Kumar et al. utilized Rluc bound to an oligonucleotide probe and a QD on the nucleic acid target [322]. Hybridization of the probe to target strands increased resonance energy transfer and QD emission. Cell-based BRET assays as well as in vivo BRET imaging are also becoming more popular as more BRET pairs become available for the selected applications as described, although care should be taken if quantitative BRET data is desired [285,287,314]. 6.4.7 Enzyme-Generated Chemiluminescence Conceptually, there is little difference between the mechanism of BL (Section 6.4.6) and enzyme-generated chemiluminescence (CL), other than in CL the luminophore is a synthetic substrate [278]. CL substrates include luminol and its derivatives, 1,2-dioxetanes and acridinium esters, which are brought to an excited state through an enzymatically catalyzed reaction [282]. In direct CL, enzymatic activity upon the substrate leads to electromagnetic radiation and an electronically excited intermediate, which luminesces. Indirect or sensitized CL occurs when the energy is instead donated to another molecule, which in turn luminesces [323]. Table 6.4 describes some common CL substrates, processing enzymes, and chemical reactions. Applications of CL include immunoassays, protein blotting, DNA probe assays [324], detection systems in separative and flow-assisted analytical techniques [325], measurement of target analytes with biospecific probes, measurements of substrates, cofactors, or quenchers, and in vitro and in vivo imaging [282]. CL resonance energy transfer (CRET) is a concept and laboratory technique that is widely underutilized in current research; however, enhanced substrates, altered j209 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.4 Biological Materials Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License enzymes, and NM amplification are slowly widening the range of useful applications [282,323,326,327]. The most widely used enzymes for CRET systems include horseradish peroxidase (HRP) and the HRP mimic hemin/G-quadruplex DNAzyme [327–329]. Although a variety of CL substrates exist for the HRP enzyme, luminol and its luminogenic derivatives remain the most popular (Table 6.4) [282,330,331]. As illustrated in the table, HRP oxidizes luminol to a luminescent species in the presence of hydrogen peroxide, yielding blue emission at 425 nm. Luminol is usually employed in conjunction with an enhancer such as luciferin, fluorescein, or a phenolic compound (e.g., para-iodophenol) [332–334], which is thought to increase the sensitivity of the assay through intermolecular energy transfer. HRP enzymatic conversion of acridan substrates generates higher luminescent intensity than luminol, when the luminescent acridinium esters intermediates decay, emitting yellow light (530 nm) [335–337]. Alkaline phosphatase is another enzyme that has shown promise in CRET applications. This enzyme catalyzes the oxidation of 1,2-dioxetane luminogenic substrates as shown in Table 6.4 [338,339]. CRET has shown promise in microchip electrophoresis [340], measurement of target molecules and proteins such as ATP [341], human immunoglobulin G (IgG) [342], alpha fetoprotein (a cancer marker) [343], microRNA [344], DNA, metal ions, and aptamers [328], with biospecific probes that are often single-stranded DNA. Analogous to BRET, there is no outside excitation, though generally the QY in CL is lower than BL. Much work has been done using NMs, including QDs [326], Au NPs [343], and graphene to enhance the QY, overall CRET brightness, and usability. A great review of many of these techniques can be found in Ref. [327]. In addition to NPs, magnetic beads have also been used to create a sensing platform that aids in a separation protocol when working with complex protein samples such as serum [342]. Recently, work by the Willner group has focused on using CRET to generate photocurrents. Through the use of a QD acceptor associated with electrical leads, CRET occurring through the enzymatic action of hemin/G-quadruplex HRP on luminol with triethanolamine as an electron donor, results in a detectable photocurrent [345]. CL and CRET research continues to evolve at a relatively slow pace compared to FRET and BRET. Recent advances with the use of QDs and other NMs, coupled with the large number of recombinant enzymes available, the low-cost commercial substrates, and control over emission wavelength, will drive further exploration of CRET for sensors and other applications. 6.5 Inorganic Materials Inorganic materials typically take the form of chelates, doped nano- or microparticles or NMs. They have a range of unique properties including bright luminescence, strong quenching abilities, large Stokes shifts, and long luminescent lifetimes that make them highly desirable for energy transfer applications as discussed in more detail later. j211 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 6.5.1 Luminescent Lanthanide Complexes and Doped Nano-/Microparticles Luminescent lanthanides are a prominent class of long-lifetime fluorophores used by the energy transfer research community for a wide range of bioapplications, mainly focused on in vitro and in vivo sensing and imaging [9,346–349]. The majority of the trivalent lanthanide ions are luminescent, with terbium (Tb), Eu, samarium (Sm), thulium (Tm), and dysprosium (Dy) emitting in the visible spectrum, while ytterbium (Yb), neodymium (Nd), and erbium (Er) emitting in the near-IR. Their emission spectra are comprised of several well-separated narrow lines coupled with long excited-state lifetimes on the order of several milliseconds. Long lifetime dyes (fluorescent lifetime t > 100 ns–ms) have a number of technical advantages over conventional fluorescence dyes (t ¼ 1–5 ns). The principal benefit arises from the ability to gate out (through time-resolved measurements) background fluorescence from direct excitation of acceptor dyes, scattering, and autofluorescence from cells and biomolecules, which can dramatically improve sensitivity. Use of time-based measurements may also necessitate more complex equipment than steady-state fluorimeters. However, because these are long-lifetime dyes (microsecond–millisecond) many standard microtiter well plate readers are available with measurement capabilities in this timescale. For bioapplications, lanthanide ions are either complexed within an organic chelate/cryptate ligand producing classical coordination metal complexes [luminescent lanthanide complexes (LLCs)] or doped into ceramic-type materials and formulated as nano-/microparticles. These complexes help to improve the optical properties of the lanthanides as well as their photostability and chemical stability, discussed in more detail later [9,347–349]. The LLC chelate ligands vary in form, but include derivatives of polyaminocarboxylates, cyclen, hydroxyquinoline, salicylamide, and phenylporphyrin, which have been recently reviewed [348]. The chelate ligands fulfill a number of functional roles in the development of successful LLC bioprobes, with ongoing research to further improve/match their properties to target applications [3,347–351]. First, the lanthanide ion must be tightly bound within these chelate complexes, resulting in higher thermodynamic and photochemical stability, which shields the lanthanide ion from the quenching effects of the surrounding solution. Second, compared to common dyes, lanthanide ions have very low extinction coefficients (1 M1 cm1), making them difficult to excite directly. Thus, the chelate label contains an organic chromophore, referred to as the lightharvesting antenna molecule or sensitizer, which is placed in close proximity to the ion. The sensitizing molecule absorbs incident light and due to close proximity transfers this energy to the lanthanide ion, presumably by a Dexter mechanism. Finally, the chelate label should possess a reactive group allowing bioconjugation. Commercial sources of lanthanide probes include CIS-Bio International (cryptate-based probes), PerkinElmer (LANCE1), Life Technologies (LanthaScreenTM), GE Healthcare (europium–TMT chelates), Lumiphore, and Sigma-Aldrich. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 212 Other than their long luminescent lifetime, LLCs have a number of other properties that make them excellent donors in luminescent resonance energy transfer (LRET) studies. For example, LLCs generally have a large effective Stokes shift, due to the separation of the absorption (by the organic chelate) and the emission spectra of the complex (from the lanthanide ion following energy transfer from the organic chelate), which reduces the potential for direct excitation of the acceptor. In addition, their emission spectrum manifest in the form of multiple, distinct, and sharp emission bands, allowing LLCs to be coupled to multiple acceptor dyes [352,353]. Tb, for example, has good spectral overlap with fluorescein, rhodamine, Cy3, and a number of AlexaFluor dye acceptors (Figure 6.19a) [346,352,353]. LLCs have also been found to undergo a phenomenon termed nonoverlapping FRET (nFRET), a mechanism not fully understood, which occurs between a LLC donor and a spectrally nonoverlapping acceptor [354,355]. Using a DNA hybridization assay, Vuojola et al. investigated nFRET between Eu(III) chelate and various AlexaFluor dyes (with varying degrees of spectral overlap with the Eu donor). They found nFRET to be very efficient over short distances, more efficient than predicted using conventional FRET theory, and unlike FRET, nFRET was found to be temperature dependent, leading the authors to conclude that a thermal excitation process was involved as part of the nFRET mechanism [354]. LRET using LLCs has been applied in a number of applications, including monitoring protein–protein interactions in cells [356], monitoring orthogonal ligand-dependent protein–peptide binding events [352], high-throughput screening of potential drug candidates [357], and numerous in vitro bioassays [347,353,358–361]. Kupstat et al., for example, developed a homogeneous time-resolved immunoassay for prostate-specific antigen (PSA) that was sensitive and quantitative, and could be incorporated into a point-of-care testing (POCT) device [360]. A sandwich immunoassay format was used in which the two antibody species that recognized and bound to different epitopes on the PSA were labeled with either the donor (Eu trisbipyridine) or the acceptor (APC protein). The presence of PSA brought the two antibodies and hence the donor/acceptor species into close proximity, resulting in LRET, with LODs two orders of magnitude below the clinical PSA cutoff of 4 ng/ ml. A similar format, using a Tb donor and five different acceptor dyes, was recently used to detect five different lung cancer tumor markers simultaneously in a 50 ml human serum sample [362]. Li et al. developed an adenosine sensor using an aptamer-based sensor design, which functioned by inducing a conformational change that disrupted the LRET (Figure 6.19b) [361]. The sensor was able to detect selectively 60 mM of adenosine in undiluted serum samples. The marriage of LLC donors and QD acceptors is a powerful combination in LRET studies and takes full advantage of the many unique properties each brings to the table, such as bright fluorescence (QDs), large Stokes shifts (both the QDs and the LLCs), and time-gated measurements (LLCs) [363]. This donor/acceptor combination has found application in luminescent microscopy, time-resolved immunoassays, measuring protease activity, and detecting nucleic acid hybridization [364–367]. Algar and coworkers in particular have developed a series of time-gated FRETrelays, demonstrating the use of QDs as simultaneous acceptors and donors in j213 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 214 bioassays for monitoring protease activity and nucleic acid hybridization [365,366]. In the case of the protease activity bioassays, the authors were able to demonstrate multiplex protease activity detection using a single QD color (Figure 6.20) [368]. Here, the QD (which acts as both a donor and an acceptor) was functionalized with peptide substrates for trypsin (labeled with a luminescent Tb complex donor) and chymotrypsin (labeled with an AlexaFluor dye acceptor). Activity of chymotrypsin resulted in a decrease in prompt FRET (between the QD and the AlexaFluor dye), while trypsin activity resulted in loss of the time-gated FRET (between the Tb complex and the QD). LLCs have also been incorporated into thin-film layers or NPs, either within the core–shell structure or bound to ligands on the NP surface, for a range of applications, including LRET-based bioassays, molecular imaging, and multiplex signal labels [369–373]. Song et al. developed core–shell nanostructures aimed at improving the LRET efficiency of the NMs, which comprised a rhodaminefunctionalized silicon dioxide (SiO2) core surrounded by a Tb chelate-modified SiO2 shell [372]. The resulting core–shell nanostructures had an energy transfer efficiency of 80%, a large F€orster distance range of 5.7–11.3 nm, and an emission lifetime of 0.25 ms. Besides LLCs, lanthanide ions are also doped into host ceramic materials, such as oxides and fluorides, to generate phosphor/luminescent materials with unique upconverting properties [upconverting phosphors (UCPs)]. Upconversion is a nonlinear phenomenon where a material sequentially absorbs long-wavelength photons and subsequently emits shorter wavelength emission, that is, converts red to visible light, a different mechanism to multiphoton absorption, where the photons are absorbed simultaneously [348,374]. The UCPs are routinely formulated as NPs (UCNPs), where they have all the benefits of the LLC materials (i.e., timeresolved measurements, sharp emission profiles, etc.), but in addition UC of the excitation light makes them excellent biolabels for in vivo imaging and in vitro bioassays. UCNPs allow the use of cheaper red excitation sources, avoid background autofluorescence from complex biological samples (improving sensitivity), and use near-IR excitation (typically 980 nm) allowing greater tissue penetration depths, while minimizing photodamage to biological samples [375–379]. The most common crystalline host material used for generating UCNPs is the fluoride NaYF4 that is either doped with one type of lanthanide species or, as is more common, codoped with two lanthanide species [374,377]. Codoping improves the UC efficiency with Yb3þ and Er3þ, representing a popular combination. 3 Figure 6.19 LLC materials for FRET. (a) Excitation (Ex) and emission (Em) spectra of Tb3þ chelate (black) versus fluorescein (green), and Alexa633 (red). (Reprinted with permission from Ref. [352]. Copyright 2007, American Chemical Society.) (b) Scheme of the adenosine sensor design based on a Tb complex conjugated to a DNA aptamer. (c) Steady-state emission spectra of the aptamer sensor upon the addition of increased concentrations of adenosine in the HEPES buffer solution (lex ¼ 344 nm). (d) Emission intensity of the sensor at 545 nm as a function of adenosine concentration. Inset: Shows the selectivity of the sensor toward adenosine over other nucleosides at 5 mM concentration. (Reprinted with permission from Ref. [361]. Copyright 2012, American Chemical Society.) j215 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.20 LLC–QD combinations. (a) Principle of the time-gated Tb ! QD ! A647 FRET relay. Optical excitation of the conjugates yields excited-state Tb and an excited-state QD ( ), and FRET2 is observed on a prompt timescale (emission <100 ns and integration time 20 ms). The extent of FRET2 is measured via the ratio of prompt A647 and QD PL, rp. Following a suitable time delay (60 ms) during which the QD returns to its ground state, FRET1 and subsequent FRET2 can be observed. The extent of FRET1 is measured in a time-gated observation window via the ratio of gated QD and Tb PL, rg. (b) Schematic of a time- gated FRET relay for multiplexed protease sensing. A central CdSe/ZnS QD is coated with compact zwitterionic ligands (CL4) and assembled with polyhistidine (His6)appended peptide substrates. The peptides, labeled with either Tb or A647, serve as substrates, SubTRP and SubChT, for TRP and ChT, respectively. The cleavage sites are highlighted in the peptide sequences. Proteolytic activity disengages FRET and alters the prompt and gated PL ratios, qp and qg, which are used as analytical signals. (Reprinted with permission from Ref. [365]. Copyright 2012, American Chemical Society.) Here, the Yb3þ dopant acts as the near-IR absorbing ion (sensitizer), while the Er3þ acts as the emitter/activator ion [374]. UCNPs can be prepared using a number of synthetic procedures, including precipitation/coprecipitation, hydrothermal/solvothermal thermolysis-based techniques, and laser annealing; for more detail refer to recent publications [377–384]. Ultimately, the goal is to prepare highly crystalline structures (which improves the overall UC efficiency) that have a small particle size combined with low size distribution, uniform dissemination of the doped lanthanide ions, good aqueous solubility, and the ability to bioconjugate, if required Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 216 [374,377,380,385,386]. Common techniques to improve the aqueous solubility of these inherently hydrophobic materials include polymer surface coatings (such as silica shells), layer-by-layer techniques, ligand exchange, and surfactant addition (including phospholipids), reviewed in Refs [376,377,380]. Lanthanide materials doped into oxides and fluorides, which take advantage of time-resolved measurements (but not the UC potential), have been used in LRET studies for a range of applications including labels in bioassays, imaging, and NP/ bioconjugation characterization [387–390]. However, fully exploiting the UC properties is more common and UCNPs have a wide range of applications, including display devices, solar cells, optoelectronics, lasers, catalysis, and biolabeling, such as bioassays, imaging, and therapy, which have been the topic of recent reviews [374,377,380,385,386,391,392]. In most cases the UCNPs donors are coupled with the traditional organic fluorophore acceptors, although there have been reports of UCNPs coupling with QD [393] or Au NP [394] acceptor NMs. LRET-based in vitro sensors using UCNPs have been successfully developed for small molecules, such as ammonia [395] and glucose [394], nucleic acid hybridization [396], and protein detection, such as caspase-3 [397] and avidin [392]. UCNPs combined with energy transfer assays have also been used in various imaging studies, for example, to look at the intracellular fate of small interference RNA (siRNA) upon uptake into cells [398]. Cheng et al. developed multicolor UCNPs for multiplex in vivo mouse imaging by tuning their emissions via LRET [399]. NaYF4 NPs doped with Er3þ/Yb3þ (green emission – donor) were functionalized with an amphiphilic polymer before being loaded with the fluorophores rhodamine B, rhodamine 6 G, or the quencher Tide Quencher 1 (acceptor) to produce three different colored UCNPs (Figure 6.21). To demonstrate the imaging utility of these materials, five UCNP materials – the three described plus unmodified NaYF4: Yb, Er (green) and NaYF4: Yb, Tm (red) – were injected subcutaneously into the back of nude mice, excited using a 980 nm laser and imaged using a MaestroTM imaging system (PerkinElmer), which captured multispectral fluorescence images (Figure 6.21). Spectral unmixing by the Maestro imaging system clearly distinguished where each population of the UCNPs was located in the mouse model, demonstrating the multiplex capability of the LRETcolor tuned UCNPs, when combined with the spectral imaging technology. 6.5.2 Luminescent Transition Metal Complexes Transition metals integrated into organic complexes, either as classical coordination metal complexes or as organometallic compounds (metal complexes containing at least one metal–carbon bond), are found to have unique luminescent properties that typically arise from a triplet metal–ligand charge transfer process and are reviewed in Refs [9,400–404]. These complexes possess a number of favorable characteristics that make them suitable for luminescent applications, including long excited-state lifetimes (100 ns–ms), high photostability, and often a large Stokes shift. These complexes have found particular application in cell imaging [9,400–404]. Of the transition metals, ruthenium (Ru) complexes remain the most popular, however j217 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.21 Multicolor imaging of UCNPs. (a) Multicolor UCL images of three UCNP– dye complexes and a mixture of the three. The images were obtained by the Maestro in vivo imaging system after spectral unmixing. (b) UCL emission spectra of solutions of UCNP1, UCNP2, UCNP1/RhB, UCNP1/R6 G, and UCNP1/TQ1 recorded by the Maestro in vivo imaging system under the 980 nm NIR laser excitation. Inset: Fluorescence spectra of RhB and R6 G under green light excitation. (c) Multicolor in vivo UCL imaging of LRET-tuned UCNPs in mice. Left image: In vivo multicolor UCL images of a nude mouse subcutaneously injected with five colors of UCNPs solutions after spectral unmixing. Right image: A white light image of the imaged mouse. (Reprinted with permission from Ref. [399]. Copyright 2011, American Chemical Society.) increasingly iridium (Ir), rhenium (Re), and occasionally osmium (Os) and platinum (Pt) have been used in cell imaging applications [9,400–405]. The current factor limiting widespread adoption of these types of materials is probably commercial availability, as most of the materials are synthesized in-house. Sigma-Aldrich offers a series of reactive Ru complexes, originally developed by Lakowicz as anisotropy labels [406,407]. These Ru complexes have lifetimes of t 500 ns, small extinction coefficients (14 500 M1 cm1), relatively low QYs (0.05), high photostability, fairly large Stokes shift, and absorption close to the visible spectrum, but, most importantly, they are functionalized with moieties that facilitate bioconjugation. While cell imaging dominates the biological applications of these luminescent transition metal complexes, there have been some examples of their use in LRETbased studies. Ru complexes have been used as donors in immunoassays for human serum albumin (HSA) [407,408] and for CO2 when coupled with the environmentally sensitive Sudan III disazo acceptor dye [409]. There are also examples of Ru Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 218 complex acceptors, in metal complex – protein binding studies and thrombin activity assays [410,411]. More recently, Ru complexes have played the role of both acceptors and donors in a three-color FRET relay for studying DNA and polypeptide dynamics and interactions [412,413]. Os(II) complexes have been coupled as acceptors in FRET studies to QD donors [414,415], and Ir(III) FRET probes have been developed for cysteine and homocysteine detection [416]. A Zn porphyrin cofactor in Zn(II)substituted horse heart cytochrome c was shown to serve as a donor to a Alex660 acceptor in cytochrome c unfolding studies [417]. CPEs (see Section 6.3.5), modified with transition metal complexes, have also been developed for protein sensing applications [138,418]. The CPEs are composed of main donor segments and transition metal complex-modified acceptor units, which in aqueous solution undergo polymer aggregation, resulting in efficient FRET. Addition of certain proteins disrupts the polymer aggregate structure, resulting in a measurable decrease in FRET efficiency. Demonstrations include HSA detection using Pt(II)modified CPEs [418] and histone detection using Ir(III)-modified CPEs [138]. 6.5.3 Noble Metal Nanomaterials (Gold, Silver, and Copper) Au, Ag, and other noble metal, such as Cu, NMs exhibit unique size- and shapedependent optical properties, due to surface plasmon resonances in the visible range (see Figure 6.22 for Au example) [419,420]. These particles typically have larger extinction coefficients (105 cm1 M1), more stable/nonfluctuating signal intensities, and greater resistance to photobleaching when compared to small-molecule fluorophores [421]. Due to their strong absorbance, they are often used as quenchers Figure 6.22 Au NPs. (a) Normalized UV–Vis absorption spectra of Au NPs with different sizes. (b) Photographs of the colloidal Au NPs with different diameters (2.4–89 nm). Concentrations of Au NPs are 670 nM (2.4 nm), 56 nM (5.5 nm), 17 nM (8.2 nm), 2.3 nM (16 nm), 0.17 nM (38 nm), and 0.013 nM (89 nm). (c) TEM images of Au NPs. Average size and standard deviation are reported for each sample. Scale bars are 20 nm (top) and 40 nm (bottom). (Reprinted with permission from Ref. [419]. Copyright 2011, American Chemical Society.) j219 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations (acceptors) in FRET applications; however, highly luminescent Au and Ag QDs or nanoclusters have been synthesized, leading to their potential role as fluorescent FRET donors/acceptors [422–424]. Excellent reviews on the properties of Au [425,426] and Ag [427–429] NMs commonly used in optical applications can be found in the literature. Au NPs can be produced in various sizes using either the citrate reduction (16– 147 nm diameter) or the Brust–Schiffrin method (1.5–5.2 nm diameter) [425,430]. These NPs can also be produced in various shapes such as spheres, rods, shells, cages, and plates [431–433]. Similarly, Ag NPs are often produced in various sizes through the reduction of Ag salts, typically a stronger reducing agent leads to smaller sized particles (<40 nm), which are often less stable over time than the larger NPs [434]. Biogenically produced nanoAg provides an environmentally friendly synthesis route [435], and various oligonucleotide sequences have been used to template Ag nanoclusters [422]. Commercial sources of Au and Ag NMs are available in a wide variety of geometries such as spheres, rods, and shells from relatively newer manufacturers such as nanoComposix, NANOCS, and NanoPartz. More specialized companies such as NanoRod, LLC or Microspheres-Nanospheres offer a wide variety of Au nanorods or metallic nanospheres, respectively. Commercially available copper NPs are less common but can be purchased in organic solution from SkySpring Nanomaterials, Inc. One of the intrinsic benefits of using Au and Ag NPs is that they are readily functionalized with ligands containing specific terminal chemical moieties (e.g., carboxyl or amine) or biomolecules through reactions with exposed thiol groups that directly attach to the NM surface via formation of an Au–sulfur (S) [436] or AgS bond [437]. Companies such as Structure Probe Inc., Nanoprobes, EB Sciences, and Research Diagnostics Inc., offer an extensive array of colloidal Au in many sizes, which are available functionalized with a variety of bioconjugates. British Biocell International offers a variety of colloidal Au and Ag also prefunctionalized as chemical or biological conjugates. Au is by far the most commonly used material in optical applications, and while Ag materials are mostly utilized for their antimicrobial properties, the increased use of Ag NMs in optical applications stems from their high QYs, photostability, and strong fluorescence intensities [427,438–443]. As a result of their native oxide layer, copper NMs have been observed to enhance fluorescence signal [444], however, recent syntheses that decrease the native oxide layer may increase the use of Cu NPs as fluorescence quenchers in future optical applications [445]. In terms of energy transfer studies, the interaction of noble metal NMs and fluorophores can be quite complex, resulting in either quenching or plasmonic enhancement of the proximal fluorophores fluorescent signal [446]. Plasmonic enhancement, observed in metal NPs coated with fluorescent dyes, is an energy transfer-type phenomenon between the excited-state fluorophore and the plasmon resonance of the proximal metal surface/particle [428,429,447–449]. Successful plasmon enhancement requires careful spacing between the fluorophore and the metal structure and factors such as metal type, NP size, and fluorophore can all influence this complex process [450–456]. Plasmon enhancement has been Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 220 exploited to increase FRET efficiency between DNA bound fluorophores [453–456], and more viable configurations are expected in the near future. Au and Ag (to a lesser extent) NMs are more commonly used for their “superquencher” abilities, that is, the ability to quench fluorescence over a broad range of wavelengths. The fluorescence quenching ability is dependent on a number of factors, including the NP size and shape, fluorophore distance from the NP surface, fluorophore dipole orientation, and the amount of overlap between fluorophore emission and NP absorption spectra, all of which influence the radiative and nonradiative decay rates [432,445,457–459]. A number of detailed studies have characterized the fluorescence quenching, via dipole–metal interactions, of various fluorophores attached near the surface of Au NPs, and both FRET and nanometal surface energy transfer (NSET) mechanisms have been proposed for emission quenching, which have a d6 and d4 distance dependency on efficiency, respectively [445,460–465]. The NSET model extends the efficient nonradiative quenching distance between the fluorophore and the proximal Au NP, effectively extending the molecular ruler capabilities, and this has been put to good use in a number of energy transfer studies [445,460–465]. The quenching abilities of Au NMs in FRET/NSET configurations are commonly exploited in a number of bioassay formats. Molecular beacon-based assays for DNA sensing measurements, for example, produce 100-fold sensitivity enhancements using Au NMs compared to previous dye–dye combinations [466–469]. Au NP–fluorophore complexes show promise for in vitro diagnostic applications as “noses,” with the ability to discern between various bacteria species and strains, proteins in complex solutions, and between cancerous and healthy cell lines [470]. These complexes have also been used for the detection of malaria antigens [471], DNA analysis [472,473], and for in vivo probes for reactive oxygen species, hyaluronidase, and protease detection systems [458]. Aptamers modified with fluorophores and subsequently combined with Au NPs have been used for the multiplex detection of adenosine, Kþ ions and cocaine [474], and nanorulers for measuring binding-site distances on live cell surfaces [475]. In addition, Au NPs have been tested as quenchers for semiconductor QDs (see Section 6.5.5). QD–Au NP systems have been used as probes to monitor real-time intracellular gene expression [476], to detect DNA hybridization events [477–480], and for TNT detection [481]. Polystyrene microspheres surface modified with Au NPs and QDs have been proposed as suitable FRET-probes for bioassays, including measuring protease activity [482]. Results from these Au NP–QD quenching demonstrations suggest that this FRET configuration has tremendous potential. Besides the lower background and improved sensitivity, the ability to label both the Au NP and QD with multiple biological moieties may improve avidity. While Au and Ag NPs are typically used for their quenching abilities, other applications utilizing their ability to scatter or fluoresce light are increasing. Highly fluorescent Au and Ag QDs consisting of only a few clusters of noble metal atoms have been synthesized [423,424], and they show potential as labels in in vitro and in vivo imaging [483]. Larger clusters of a few nanometer thicknesses show promise j221 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations as low photobleaching alternatives in cancer cell imaging [484]. Much like their semiconductor counterparts, these Au and Ag QDs (or nanodots, NDs) have sizetunable emission maxima, shifting to longer wavelengths with increasing nanocluster size, although emission can also be influenced by other factors such as surface ligands and crystallinity of the NM [424]. Although not strictly energy transfer based, plasmonic rulers consisting of two plasmonic particles in close proximity, capable of detecting molecular binding and folding events, offer several advantages over traditional dye-based FRET applications in that they prevent photobleaching as well as allow for a 10-fold increase in measurement range [485,486]. Plasmonic rulers may also be fabricated in three dimensions allowing a more spatially complete understanding of complex molecular events [487]. The wide variety of materials and applications incorporating the light-altering properties of noble metal NPs suggests the increasing importance of these particles in optical applications. 6.5.4 Silicon-Based Materials A number of Si-based materials, considered metalloid in nature, have been found to have intrinsic fluorescent properties that are worth mentioning. Silole molecules and polymers, for example, are Si-containing five-membered cyclic dienes structures that are found to become highly fluorescent upon aggregation [488]. The silole molecule 1,1,2,3,4,5-hexaphenylsilole (HPS) generates a strong aggregation-induced luminescence at 495 nm. Amorphous silica (SiO2) NPs have also been found to exhibit inherent luminescence due to oxygen-stabilized defects in the SiO2 lattice [489], although it is more common to dope silica NM structures (core, core–shell, and shell structures) with organic fluorophores in the pursuit of fluorescent NMs (see Section 6.3.4). Si NPs are a much more commonly utilized luminescent form and are increasingly used as bioimaging agents due to their low toxicity, resistance to photobleaching as well as their bright size-dependent photoluminescence and broad excitation spectra [490–495]. Si NPs have been investigated as fluorescent tags for DNA [496], photonic barcode devices [497], potentially nontoxic materials for in vivo and in vitro imaging [498] including biodegradable imaging systems [499], theranostic systems in which particles can image as well as potentially treat cancer cells photodynamically [500,501], and biomodal imaging systems in which iron-doped Si particles exhibit magnetic as well as fluorescent properties [502]. The synthesis of Si NPs remains tricky, but new methods for synthesizing and stabilizing them have been reported [503–505], including those that provide a variety of particle geometries such as “flowerlike” polyhedron [506], nanowires, and clusters [507]. An extensive review of Si NP synthesis and physical properties can be found in Refs [495,508]. Future applications using Si-based NPs as FRET donors can be expected because of the incredible photostability, tunability, and facile surface modification of these materials. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 222 6.5.5 Semiconductor Nanocrystals Colloidal luminescent semiconductor nanocrystals, or QDs, have a range of potential applications, including optoelectronics (lighting and advanced displays), optics (lasers), solar energy, biotechnology, and medicine [509]. However, since their inception as biological labels, applications in the biological arena have developed exponentially [283,510–515]. QDs can function either as a passive fluorescent label or in a more integrated/active role, where the QD both acts as a scaffold for biorecognition and is intimately involved in signal transduction, through mechanisms such as FRET, BRET, charge transfer (CT), or CRET [283,512,513,516–518]. QDs possess a number of unique electro-optical properties that make them ideal energy transfer labels [283,512,513,515,518,519]. Benefits include a size and chemical composition dependent emission that is narrow and symmetric in profile, high quantum yields and extinction coefficients (enabling single-molecule detection), broad absorption profile, large Stokes shift, large two-photon absorption properties, and excellent resistance to photobleaching and chemical degradation [283,512,513,515,518–521]. QDs have been synthesized from a range of binary and ternary alloys such as ZnS, CdSe, CdTe, InP, GaN, PbS, ZnO, InGaAs, and CdZnS, and their exact emission, which can span from UV-Vis to infrared, is found to be dependent on both the chemical composition and NP size, resulting in a tunable emission profile (Figure 6.23) [515,522,523]. For FRET, in particular, this means that QD donors can be size-tuned or “dialed in” to have better spectral overlap with a particular acceptor dye, improving FRET efficiency. The broad absorption spectra and large Stokes shift found in QDs is also of benefit for FRET studies, as it allows excitation of mixed QD donor populations at one wavelength far removed from their emissions and also facilitates selection of an excitation wavelength that corresponds to the acceptors absorption minima, thus reducing direct excitation background signals. The ability to excite multiple QD donors using a single excitation wavelength, combined with their narrow and symmetric emission (which makes deconvolution of multiple fluorescent signals simpler), makes them attractive labels for multiplex applications [32,321,513,524–529]. As with any of the potential FRET labels discussed in this chapter, there are issues to consider before using QDs. QDs have been found to blink under continuous excitation, which may be problematic for single-molecule studies. QDs have a finite size that can be both a benefit and a liability for FRET. In addition, the “as synthesized” semiconducting nanocrystals are inherently hydrophobic, requiring some type of modification, that is, surface coating, to facilitate water solubility while maintaining their optical properties. Aqueous solubility can be achieved using three main approaches: direct aqueous synthesis using hydrophilic stabilizing agents, cap exchange of hydrophobic surfactants with hydrophilic ligands, and encapsulation/ coating with amphiphilic species, polymers, or silica coatings, reviewed in Refs [512,513,515,519,530]. Encapsulation, in particular, can significantly influence the hydrodynamic radius of the final QD and hence impact the distance-dependent FRET efficiency [512,515,518,519]. A number of researchers have focused on j223 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.23 Dependence of fluorescence emission wavelengths of QDs on their chemical composition. The CdSe group is expanded to demonstrate the sizedependent absorption and emission profiles of the CdSe QDs, obtained by Peng synthesis and different heating times: 3, 5, 7, 10, 14, 20, 25, and 30 min. (Reprinted with permission from Ref. [515]. Copyright 2013, Elsevier.) developing ligands that can facilitate both aqueous solubility and subsequent bioconjugation of CdSe/ZnS QDs, while keeping the overall hydrodynamic radius compact [512,513,515,519,531]. These ligands have evolved from simple designs that facilitate aqueous solubility to multifunctional modular designs that comprise an anchor group that interacts with the QD surface (commonly a monodentate or Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 224 bidentate thiol), a hydrophilic segment that imparts solubility (e.g., PEG, zwitterionic nature), and a terminal functional group that can provide solubility and be used for bioconjugation (e.g., OH, NH2, COOH, Biotin, and azide or alkyne for click chemistry) [518,531,532]. Susumu et al., for example, designed compact zwitterionic inspired ligands for preparing QDs and Au NPs [533]. The zwitterionic nature of the ligands greatly improved the pH stability of the QDs, compared to dihydrolipoic acid (DHLA) only preparations, without impacting the hydrodynamic radius of the functionalized QDs (Figure 6.24) [533]. In situations where FRET efficiency is found to be low, the finite QD size can be an advantage allowing multiple acceptors to be assembled around the central QD scaffold, improving the FRET efficiency and potentially enhancing subsequent biomolecular interactions (as recently demonstrated in proteolytic digestion studies [368]), but this may not be desirable for all applications [3,518]. In addition, dimensionality of the NM may be a factor influencing the FRET efficiency, as found in the case of a QD donor (spheres versus rods) coupled to multiple acceptors [534–536]. The broad absorption profile of QDs serves them well as FRET donors, however it can be problematic for their application as FRET acceptors, where direct excitation of the QD by the excitation source is an issue [517]. With careful choice of the donor species, there are a number of instances where QDs are excellent energy transfer acceptors, including BRET and CRET assays (Sections 6.4.6 and 6.4.7, respectively), which do not require an excitation source, or the use of long-lifetime donors, therefore allowing the performance of time-gated/resolved measurements, which can factor out any direct QD excitation [517,524]. There are a number of commercial suppliers of QD materials, recently reviewed in Ref. [515], that are either functionalized with chemical handles that facilitate bioconjugation (e.g., OH, NH2, COOH, and -biotin) or are bioconjugated with SA or various secondary antibodies (e.g., goat antihuman, goat antimouse, or goat antirabbit). While there are increasing varieties of commercially available QD materials, the most popular ones still remain the CdSe and CdTe core materials and the CdSe/ZnS core–shell QDs. There are reviews and detailed monographs describing QDs synthesis using various materials, generally involving wet chemistry techniques such as high-temperature organometallic synthesis, microwave or gamma irradiation, and aqueous colloidal and sol–gel methods [515,523,537,538]. Various biotemplated fabrication approaches have also been proposed for QD production, ranging from whole organism synthesis (bacteria, yeast, and viruses) to biomolecule-based ligands that facilitate nucleation and capping of the QDs during synthesis (nucleic acids and peptide sequences) [539–543]. While greener in terms of reagents and reaction conditions, these biofabrication methods tend to produce lower quality QDs, in terms of QY and size polydispersity, compared to high-temperature chemical synthetic techniques, but this may improve as our understanding of the underlying mechanisms that govern these syntheses grows [539–543]. As discussed in Section 6.2, there are a number of diverse strategies that exist for attaching biomolecules to QDs, including covalent coupling, electrostatic/metal j225 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.24 Improving aqueous solubility of QDs through DHLA surface ligands. (a) Chemical structures of DHLA and the DHLAbased ligands used to stabilize QDs. (b–d) Physical characterization of a series of compact ligand (CL)-coated QDs. (b) Gel electrophoretic separation of 550 nm emitting QDs capexchanged with the indicated ligands. Gels were run on 1% agarose gel in 1 TBE buffer (pH 8.3) at 7 V/cm for 10 min. (c) Hydrodynamic size distribution of 550 nm emitting QDs capexchanged with DHLA: 10.8 (2.7 nm), CL1: 8.6 (1.8 nm), CL2: 9.3 (1.7 nm), CL3: 9.5 (2.1 nm), CL4: 9.8 (2.2 nm), and DHLA-PEG750-OCH3: 11.5 (2.5 nm) measured by dynamic light scattering. Data is plotted in arbitrary units of scattering intensity. (d) PL images (left) for a set of 0.5 mM QDs capped with the CL1 compact ligands in different buffers at pH 2–13. The 550 nm emitting CdSe/ZnS QDs were used and excited with a UV lamp at 365 nm. Images were taken <20 min and 4 weeks after sample preparation. PL images (right) for a set of 0.5 mM–550 nm emitting QDs coated with DHLA or the indicated compact ligands in 3 M NaCl solution. Images were taken 1 day after sample preparation for DHLA and after 90 days for CL1CL4. (Reprinted with permission from Ref. [533]. Copyright 2011, American Chemical Society.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 226 affinity-driven self-assembly, and biotin–avidin chemistry (Figure 6.25) [283]. However, given their NM nature, additional issues (also discussed in Section 6.2), such as the influence of the surface ligand on bioconjugation reactions and subsequent biomolecular interactions and how the attachment chemistry influences the NM stability and biomolecule orientation, should be considered for QDs [30,33–36,515]. QDs as donors in FRET applications have been coupled to a range of acceptor materials for energy transfer studies, including fluorescent dyes, NMs (such as Au and carbon), fluorescent proteins, and polymers (Figure 6.26a) [79,184,283,462,482,513,516,517,520,524,544–548]. As mentioned earlier, they are also adept FRET acceptors when coupled with the appropriate donors, such Figure 6.25 QD bioconjugation – an illustration of some selected surface chemistries and conjugation strategies that are applied to QDs. The gray periphery around the QD represents a general coating. This coating can be associated with the surface of the QD via (e) hydrophobic interactions, or ligand coordination. Examples of the latter include (a) monodentate or bidentate thiols, (b) imidazole, polyimidazole (e.g., polyhistidine), or dithiocarbamate (not shown) groups. The exterior of the coating mediates aqueous solubility by the display of (c) amine or carboxyl groups, or (d) functionalized PEG. Common strategies for bioconjugation include (a) thiol modifications or (b) polyhistidine or metallothionein (not shown) tags that penetrate the coating and interact with the surface of the QD, (f) electrostatic association with the coating, (g) nickel-mediated assembly of polyhistidine to carboxyl coatings, (h) maleimide activation and coupling, (i) active ester formation and coupling, and (j) biotin labeling and SA–QD conjugates. The figure is not to scale. (Reprinted with permission from Ref. [283]. Copyright 2010, Elsevier.) j227 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 228 as long-lifetime lanthanide materials (Figure 6.26b) [513,517,549]. QD donors/ acceptors have also been applied in a number of studies looking at multi-FRET processes, discussed in more detail in Section 6.6. Algar et al., in addition, recently demonstrated QDs as simultaneous acceptors and donors in time-gated FRET relays bioassays for monitoring protease activity and nucleic acid hybridization (Figure 6.20) [366]. QDs have also been shown to be sensitive to CT processes and coupling with redox active species, such as dopamine, Ru, rhodamine B, and Os, which often leads to a competition between energy and charge transfer processes in the resulting complex (Figure 6.26c) [415,513,550–553]. Since our previous review [3], the use of QDs in energy transfer-based applications has truly blossomed and has been reviewed in a number of excellent monographs [283,513,515–517,524,554,555]. In vitro applications dominate the QD FRET literature, including assays for detection of specific targets ranging from small molecules to proteins and whole organisms (e.g., bacteria), enzyme activity monitoring, tracking intracellular gene delivery, solid-phase assays, and QD-enabled singlemolecule detection. There are hosts of biomolecular interactions/processes that are studied using QD energy transfer formats and these include nucleic acid interactions, binding protein conformation changes, antibody binding, aptamer interactions, and protease cleavage [283,513,515–518,524,554]. Assay formats found in conjunction with QD FRET are quite varied and include (i) cleavage-based assays (e. g., proteases, kinases, and DNAzymes) (Figure 6.27) [321,365,368,526–529,556– 558], (ii) conformational change-based assays (e.g., aptamers, binding proteins, and DNA molecular beacons) [518,559–561], (iii) displacement assays (e.g., antibodies and binding proteins) (Figure 6.28) [554,562], (iv) various immunoassays (including direct, displacement, and sandwich) [367,562–564], (v) nucleic acid hybridization [518,555], and (vi) assays based on acceptor spectral changes (mainly pH or ion sensitive dyes) (Figure 6.16) [79,527,556]. Given the unique photophysical properties of QDs, our increasing fundamental understanding of these unique materials when used in energy transfer configurations, and the availability of improved synthesis and bioconjugation methods, we can expect continued utilization in many FRETbased biological assays, with increasing emphasis on multiplexed and in vivo detection [321,524,526–529,549,565]. 3 Figure 6.26 QDs as FRET acceptors and donors. (a) QDs are good FRET donors for fluorescent proteins (FPs), dye, and Au NP acceptors. The dashed circle represents an arbitrary F€ orster distance (R0) measured from the QD center. The scale on the right indicates how R0 proportionally increases as the number of proximal acceptors (a) increases. Conversely, QDs can function as acceptors for Tb complexes and BL luciferase donors. (b) Qualitative spectral overlap (shaded) for a 625 nm emitting CdSe/ZnS QD as (i) donor to fluorescent dye acceptor (Alexa647, A647) and (ii) acceptor to a Tb chelate donor. (c) CT quenching is an alternative method of modulating QD PL: (i) an electron acceptor (e. g., quinone) has an unoccupied energy level intermediate in energy to the 1Sh and 1Se bandedge states to which the excited QD transfers an electron, and (ii) an electron donor (e.g., Ru phenanthroline) has an occupied intermediate energy level and transfers an electron to the QD. Charge transfer inhibits radiative recombination of the exciton. Both the redox active species are illustrated as peptides conjugates. (Reprinted with permission from Ref. [513]. Copyright 2011, American Chemical Society.) j229 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.5 Inorganic Materials j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.27 Cleavage-based FRET assays. (a) A protease assay where (i) an acceptor dyelabeled peptide is assembled on a QD donor via a polyhistidine tag. The QD-dye proximity in the bioconjugate is sufficient for FRET. (ii) Protease activity (scissors) cleaves the peptide and disrupts FRET, restoring the QD PL. (Reprinted with permission from Ref. [283]. Copyright 2010, Elsevier.) (b) Proteolytic assay data from exposing a constant concentration of 550 nm emitting QDs conjugated to four Texas Red substrate peptides to a constant concentration Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 230 6.6 Multi-FRET Systems Multistep FRET is a naturally occurring process exemplified by the light-harvesting systems found in many biological species, which allows them to harness light not readily absorbed by chlorophyll for photosynthesis [254,566]. Phycobilisomes are supramolecular complexes found in blue-green cyanobacteria and various algae such as glaucophytes, red, and cryptomonad [253,254]. Phycobilisomes have a variable composition that is organism dependent, but typically consists of multiple PBP subunits (see Section 6.4.4). As mentioned in Section 6.4.4, within the phycobilisome system there are four main types of PBP – PC, PE, PEC, and APC – that bind the phycobilin chromophores (e.g., PUB, PEB, and PCB) that give these proteins their intense colors [254]. The multi-FRET process and flow of energy in phycobilisomes is PE–PC–APC–photosynthetic reaction center (chlorophyll) [254]. Energy transfer efficiency in this system approaches nearly 100%, and researchers have yet to match experimentally the complexity or efficiency of this naturally selected energy harvesting system. An increasing number of biologically inspired and artificial synthetic multi-FRET systems have been developed to precisely space or orient multiple fluorophores with the goal of characterizing and mimicking the natural light-harvesting process. Such systems have potential use in solar cells, nanoscale photonic devices, and other optoelectronic applications [269,567,568]. More commonly, these multi-FRET configurations are developed as a means to extend the optical ruler and are used to elucidate biological configurations, study protein and DNA interactions, and for biosensing applications [135,210,365,366,412,413,569]. While artificial synthetic building block structures have been developed, such as perylene bisimide-calix[4]arene arrays by Hippius et al., to control the position and orientation of chromophores, biologically inspired platforms are more common [570]. DNA is perhaps the most attractive biologically inspired platform for multiFRET configurations due to (i) its predictable structure/chemistry, (ii) the inherent ability to introduce fluorophores at specific sites, (iii) the ability to hybridize multiple dye-labeled oligos to a complimentary strand, and (iv) the ability to control the orientation of the attached fluorophores [3]. DNA can be synthesized with multiple fluorophores at specific terminal or internal sites or with thiol/amine/biotin or other 3 of caspase-3 enzyme. Derived Km and Vmax values are given. An R2 of 0.98 was obtained for the fitting of the curve. (Reprinted with permission from Ref. [556]. Copyright 2010, American Chemical Society.) (c) Multiplexed assay of proteases by using QDs with different colors on a glass slide. SA-QD525, SA-QD605, and SA-QD655 were used (from left to right). Biotinylated peptide substrates for MMP-7, caspase-3, and thrombin were conjugated to the AuNPs, and then the resulting Pep-AuNPs were associated with SA-QD525, SA-QD605, and SA-QD655, respectively: (i) SA-QDs only, (ii) SA-QDs þ respective Pep-AuNPs, (iii) SAQDs þ Pep-AuNPs þ MMP-7, (iv) SAQDs þ Pep-AuNPs þ caspase-3, (v) SAQDs þ Pep-AuNPs þ thrombin, and (vi) QDs þ Pep-AuNPs þ mixture of the respective protease and its inhibitor. (Reprinted with permission from Ref. [528]. Copyright 2008, American Chemical Society.) j231 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.6 Multi-FRET Systems j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.28 Displacement format using antibody-functionalized QDs for FRET-based TNT detection. (a) Schematic of the assay. When TNB-BHQ-10 is bound to the QDTNB2-45 conjugate, QD fluorescence is quenched via FRET. As TNT is added to the assay, it competes for binding to the antibody fragment and the QD fluorescence increases following TNB-BHQ-10 release from the conjugate. (b) Results from titration of the QD-TNB2-45-TNB-BHQ-10 assembly with TNT and the indicated TNT analogues. These assemblies were constructed using 530 nm emitting QDs. Each data point is an average of three measurements, and error bars represent the standard deviation. (Reprinted with permission from Ref. [562]. Copyright 2005, American Chemical Society.) modifications allowing custom labeling. Altering donor/acceptor spacing is facile in this configuration and allows fine-tuning of FRET efficiency [569]. Such fine-tuning and control of multi-FRET using DNA constructs has been used for light harvesting and charge separation using DNA nanostructures [269,567], development of combinatorial FRET-tags for SNP detection, [571] and DNA-based photonic wires [568]. Proteins have also been used in the development of multi-FRET configurations. Maltose binding protein (MBP), for example, has been either triple labeled with FAM, tetramethylrhodamine, and Cy5, [572] or dual labeled with QD and Cy3 before binding Cy3.5-labeled b-cyclodextrin [71] in the development of maltose biosensors. Rogers et al. incorporated a multi-FRET PheCN, Trp, and 7AW system into two model protein systems: HP35 and a designed bba motif (BBA5) to study Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 232 urea-induced conformational changes [210]. The nonnatural amino acid fluorophores, PheCN and 7AW (Section 6.4.2), were incorporated by mutation of the Cand N-terminus. Such a three-color FRET system extends the working range of the molecular ruler and yields information about the relative positions between the three fluorophores. In addition to the use of the more traditional organic fluorophores for multi-FRET applications, substantial development in the areas of NMs, time-resolved fluorescent reagents, and BRET has led to material combinations, with unique advantages in multi-FRET applications. QDs NMs can act as donors and acceptors in FRET configurations as well as provide a scaffold in which to immobilize the multiple components of a multi-FRET system. Medintz et al. developed a multi-FRET maltose biosensor using QDs (initial donor) functionalized with Cy3-labeled MBP that further bound Cy3.5-labeled b-cyclodextrin, a sugar analogue that was displaced from the QD–MBP complex in the presence of maltose [71]. Boeneman et al. developed DNA photonic wires using a QD modified with an ssDNA backbone template that bound four separate short complementary sequences (labeled with different dyes) (Figure 6.29) [568]. The system was used to study the sequential FRET from the initial QD donor to Cy3/Cy5/Cy5.5/Cy7, and it was found that while the initial QD-quenching efficiency was high (80–90%), the amount of energy subsequently emitted by Cy5 and Cy5.5 was relatively low at 2.2 and 1%, respectively, with the Cy7 acting as an IR quencher. QDs can also function as simultaneous donors and acceptors in multi-FRET systems, as demonstrated in DNA hybridization and protease activity detection assays [135,365,366]. The use of long-lifetime dyes, such as Ru complexes and LLCs, can provide additional dynamic information about a system or enhance the biosensing capabilities of a platform [365,366,412,413]. The groups of Kumke and Bannwarth, for example, created a three-color FRET system for protein and DNA analysis using a carbostyril donor–Ru complex (acceptor/relay)–anthraquinone (quencher) combination [412,413]. In the case of DNA analysis, they found that the short luminescent lifetimes gave information about the rotation of the dye molecules themselves, while the long lifetimes yielded information regarding the overall dynamics within the DNA macromolecule itself [412]. Algar et al. developed and characterized multiFRET systems for monitoring DNA hybridization and protease activity [365,366]. Here, a Tb chelate–QD–Alexa647 combination was used in which the Tb chelate assumed the role of initial donor and facilitated time-gated measurements, ultimately allowing multiplexed biosensing based on a single-color QD scaffold (discussed in more detail in Sections 6.5.1 and 6.5.5, and also see Figure 6.20). SRET, the sequential combination of BRET and FRET, is a fairly recent development in the multiresonance energy transfer arena [317,318,573]. Carriba et al., for example, labeled three different membrane receptor-interacting proteins with Rluc, GFP, YFP, or DsRed, using SRET to study the heteromers that formed upon exposure to agonists [317]. Protease activity has also been monitored via SRET (Figure 6.30) [318]. Here, the peptide-based probe comprised a peptide sequence (containing the protease-specific cleavage site) flanked by a thermostable firefly luciferase that produced yellow-green BL (BRET), and a red FP labeled with a near- j233 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.6 Multi-FRET Systems j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations Figure 6.29 Multi-FRET systems using DNA templates. (a) Schematic of the configuration consisting of 530 nm QDs self-assembled with four (His)6-peptide-DNA hybridized with Cy3 in 1/Cy5 in 2/Cy5.5 in 3/Cy7 in 4. (b) Composite PL spectra from 530 nm QD donors self-assembled with unlabeled DNA, DNA with Cy3 in position-1, Cy3 in 1/Cy5 in 2, Cy3 in 1/Cy5 in 2/Cy5.5 in 3, Cy3 in 1/Cy5 in 2/Cy5.5 in 3/Cy7 in 4. (Reprinted with permission from Ref. [568]. Copyright 2010, American Chemical Society.) infrared fluorescent dye (Alexa680) (FRET). The intact peptide probe undergoes efficient SRET, resulting in acceptor emission of the near-infrared fluorescent dye (at 705 nm). Addition of the protease results in a decrease in acceptor emission, due to disruption of the SRET process. Xiong et al. developed SRET-based NPs for near-IR in vivo imaging of the lymphatic networks and vasculature of xenografted tumors in mice [573]. The NPs were composed of a fluorescent polymer, poly[2-methoxy-5((2-ethylhexyl)oxy)-p-phenylenevinylene] (MEH-PPV), which was subsequently doped with the near-IR dye (NIR775) and its surface modified with a COOH- Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 234 Figure 6.30 Multi-FRET systems – sequential BRET then FRET – for protease sensing. (a–c) Factor Xa detection using a unique sequential BRET–FRET combination. (a) The peptidebased probe comprised a peptide sequence (containing the protease-specific cleavage site) flanked by a thermostable firefly luciferase that produces yellow-green BL, and a red FP labeled with a near-infrared fluorescent dye (Alexa680). (b) When intact the peptide probe undergoes efficient BRET/FRET resulting in acceptor emission of the near-infrared fluorescent dye (at 705 nm). (c) Addition of the protease factor Xa results in a decrease in acceptor emission, due to disruption of the BRET/FRET process, as illustrated in the time course spectra. (Reprinted with permission from Ref. [318]. Copyright 2011, Elsevier.) terminated PEG polymer that facilitated Luc8 and tumor-targeting ligand (RGD peptide) bioconjugation. BRET occurred between the Luc8 and MEH-PPV in the presence of substrate coelenterazine, with sequential FRET occurring between the MEH-PPV and NIR775. The NPs demonstrated good blood circulation and tumor targeting in mice models. Although there have been fairly limited SRET demonstrations to date, given the improved sensitivity afforded by the self-illuminating nature of these SRET-based systems utility is bound to increase. j235 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 6.6 Multi-FRET Systems j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 6.7 Summary and Outlook FRET is clearly an invaluable biophysical tool for a variety of applications, ranging from fundamental studies of the structure and conformation of biological materials and the in vivo examination of biomolecular interactions, to more applied uses where FRET signal transduction has been utilized for in vitro and in vivo bioassays, for healthcare diagnostics/screening, defense, environment, and food safety. The range of new and improved donor/acceptor probe materials continues to expand to address some of the inherent complications of more traditional FRET materials, including photobleaching, spectral cross talk and direct excitation of the acceptor species. NMs, in particular, are increasingly being used as donor/acceptor probes in FRET studies due to their many unique photophysical properties and their inherent nanoscaffolding capabilities that can be used to improve FRET efficiency. Hand in hand with donor/acceptor materials development that has expanded the use of FRET has been the evolving bioconjugation techniques, especially bioorthogonal methods that facilitate greater site-specific control of the donor/acceptor labeling. With further advances in the areas of NMs and bioconjugation techniques, we anticipate FRET to become an increasingly appreciated tool in a wide range of fundamental and applied applications. Acknowledgments K.E.S acknowledges Division of Biology, FDA and MCMi, FDA for financial support. K.E.S would also like to thank Ms. S. Spindel and Dr. K. Butler for their comments and review of this chapter. The mention of commercial products, their sources, or their use in connection with material reported herein is not to be construed as actual and/or implied endorsement of such products by the Department of Health and Human Services. References 1 Jares-Erijman, E.A. and Jovin, T.M. (2003) FRET imaging. Nature Biotechnology, 21 (11), 1387–1395. 2 Sapsford, K.E., Berti, L., and Medintz, I.L. (2004) Fluorescence resonance energy transfer: concepts, applications and advances. Minerva Biotecnologica, 16 (4), 247–273. 3 Sapsford, K.E., Berti, L., and Medintz, I.L. (2006) Materials for fluorescence resonance energy transfer analysis: beyond traditional donor–acceptor combinations. Angewandte Chemie, International Edition, 45 (28), 4562–4588. 4 Huebsch, N.D. and Mooney, D.J. (2007) Fluorescent resonance energy transfer: a tool for probing molecular cell– biomaterial interactions in three dimensions. Biomaterials, 28 (15), 2424–2437. 5 Roda, A., Guardigli, M., Michelini, E., and Mirasoli, M. (2009) Nanobioanalytical luminescence: Forster-type energy transfer methods. Analytical and Bioanalytical Chemistry, 393 (1), 109–123. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 236 6 Preus, S. and Wilhelmsson, L.M. (2012) 7 8 9 10 11 12 13 14 15 16 17 Advances in quantitative FRET-based methods for studying nucleic acids. ChemBioChem, 13 (14), 1990–2001. Padilla-Parra, S. and Tramier, M. (2012) FRET microscopy in the living cell: different approaches, strengths and weaknesses. Bioessays, 34 (5), 369–376. Zhong, W. (2009) Nanomaterials in fluorescence-based biosensing. Analytical and Bioanalytical Chemistry, 394 (1), 47–59. Hoetzer, B., Medintz, I.L., and Hildebrandt, N. (2012) Fluorescence in nanobiotechnology: sophisticated fluorophores for novel applications. Small, 8 (15), 2297–2326. Sahoo, H. (2012) Fluorescent labeling techniques in biomolecules: a flashback. RSC Advances, 2 (18), 7017–7029. Chattopadhaya, S., Abu Bakar, F.B., and Yao, S.Q. (2009) Expanding the chemical biologist’s tool kit: chemical labelling strategies and its applications. Current Medicinal Chemistry, 16 (34), 4527–4543. Sletten, E.M. and Bertozzi, C.R. (2009) Bioorthogonal chemistry: fishing for selectivity in a sea of functionality. Angewandte Chemie, International Edition, 48 (38), 6974–6998. Didenko, V.V. (2001) DNA probes using fluorescence resonance energy transfer (FRET): designs and applications. Biotechniques, 31 (5), 1106–1116. Berti, L., Xie, J., Medintz, I.L., Glazer, A. N., and Mathies, R.A. (2001) Energy transfer cassettes for facile labeling of sequencing and PCR primers. Analytical Biochemistry, 292 (2), 188–197. De Lorimier, R.M., Smith, J.J., Dwyer, M. A., Looger, L.L., Sali, K.M., Paavola, C.D., Rizk, S.S., Sadigov, S., Conrad, D.W., Loew, L., and Hellinga, H.W. (2002) Construction of a fluorescent biosensor family. Protein Science, 11 (11), 2655–2675. Jager, M., Nir, E., and Weiss, S. (2006) Site-specific labeling of proteins for single-molecule FRET by combining chemical and enzymatic modification. Protein Science, 15 (3), 640–646. Crochet, A.P., Kabir, M.M., Francis, M.B., and Paavola, C.D. (2010) Site-selective dual modification of periplasmic binding 18 19 20 21 22 23 24 25 26 27 28 proteins for sensing applications. Biosensors & Bioelectronics, 26 (1), 55–61. Hermanson, G.T. (2008) Bioconjugate Techniques, 2nd edn, Academic Press. Johnson, I. and Spence, M.T.Z. (eds) (2010) Molecular Probes Handbook: A Guide to Fluorescent Probes and Labeling Technologies, 11th edn, Life Technologies. Lim, R.K.V. and Lin, Q. (2010) Bioorthogonal chemistry: recent progress and future directions. Chemical Communications, 46 (10), 1589–1600. Prescher, J.A. and Bertozzi, C.R. (2005) Chemistry in living systems. Nature Chemical Biology, 1 (1), 13–21. Campbell, R.E. (2009) Fluorescentprotein-based biosensors: modulation of energy transfer as a design principle. Analytical Chemistry, 81 (15), 5972–5979. Chen, I. and Ting, A.Y. (2005) Site-specific labeling of proteins with small molecules in live cells. Current Opinion in Biotechnology, 16 (1), 35–40. Adams, S.R., Campbell, R.E., Gross, L.A., Martin, B.R., Walkup, G.K., Yao, Y., Llopis, J., and Tsien, R.Y. (2002) New biarsenical ligands and tetracysteine motifs for protein labeling in vitro and in vivo: synthesis and biological applications. Journal of the American Chemical Society, 124 (21), 6063–6076. Scheck, R.A. and Schepartz, A. (2011) Surveying protein structure and function using bis-arsenical small molecules. Accounts of Chemical Research, 44 (9), 654–665. Soh, N. (2008) Selective chemical labeling of proteins with small fluorescent molecules based on metal-chelation methodology. Sensors, 8 (2), 1004–1024. Ojida, A., Honda, K., Shinmi, D., Kiyonaka, S., Mori, Y., and Hamachi, I. (2006) Oligo-Asp Tag/Zn(II) complex probe as a new pair for labeling and fluorescence imaging of proteins. Journal of the American Chemical Society, 128 (32), 10452–10459. Chen, I., Howarth, M., Lin, W.Y., and Ting, A.Y. (2005) Site-specific labeling of cell surface proteins with biophysical probes using biotin ligase. Nature Methods, 2 (2), 99–104. j237 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 29 Foley, T.L. and Burkart, M.D. (2007) Site- 30 31 32 33 34 35 36 specific protein modification: advances and applications. Current Opinion in Chemical Biology, 11 (1), 12–19. Algar, W.R., Prasuhn, D.E., Stewart, M.H., Jennings, T.L., Blanco-Canosa, J.B., Dawson, P.E., and Medintz, I.L. (2011) The controlled display of biomolecules on nanoparticles: a challenge suited to bioorthogonal chemistry. Bioconjugate Chemistry, 22 (5), 825–858. Sapsford, K.E., Algar, W.R., Berti, L., Boeneman Gemmill, K., Casey, B., Oh, E., Stewart, M.H., and Medintz, I.L. (2013) Functionalizing nanoparticles with biological molecules: developing chemistries that facilitate nanotechnology. Chemical Reviews, 113 (3), 1904–2074. doi: dx.doi.org/10.1021/cr300143v Hildebrandt, N. (2011) Biofunctional quantum dots: controlled conjugation for multiplexed biosensors. ACS Nano, 5 (7), 5286–5290. Dennis, A.M., Sotto, D.C., Mei, B.C., Medintz, I.L., Mattoussi, H., and Bao, G. (2010) Surface ligand effects on metalaffinity coordination to quantum dots: implications for nanoprobe self-assembly. Bioconjugate Chemistry, 21 (7), 1160–1170. Bernardin, A., Cazet, A., Guyon, L., Delannoy, P., Vinet, F., Bonnaffe, D., and Texier, I. (2010) Copper-free click chemistry for highly luminescent quantum dot conjugates: application to in vivo metabolic imaging. Bioconjugate Chemistry, 21 (4), 583–588. Sapsford, K.E., Granek, J., Deschamps, J. R., Boeneman, K., Blanco-Canosa, J.B., Dawson, P.E., Susumu, K., Stewart, M.H., and Medintz, I.L. (2011) Monitoring botulinum neurotoxin a activity with peptide-functionalized quantum dot resonance energy transfer sensors. ACS Nano, 5 (4), 2687–2699. Boeneman, K., Deschamps, J.R., Buckhout-White, S., Prasuhn, D.E., Blanco-Canosa, J.B., Dawson, P.E., Stewart, M.H., Susumu, K., Goldman, E. R., Ancona, M., and Medintz, I.L. (2010) Quantum dot DNA bioconjugates: attachment chemistry strongly influences the resulting composite architecture. ACS Nano, 4 (12), 7253–7266. 37 Lavis, L.D. and Raines, R.T. (2008) Bright 38 39 40 41 42 43 44 ideas for chemical biology. ACS Chemical Biology, 3 (3), 142–155. Dsouza, R.N., Pischel, U., and Nau, W.M. (2011) Fluorescent dyes and their supramolecular host/guest complexes with macrocycles in aqueous solution. Chemical Reviews, 111 (12), 7941–7980. Wetzl, B., Gruber, M., Oswald, B., Durkop, A., Weidgans, B., Probst, M., and Wolfbeis, O.S. (2003) Set of fluorochromophores in the wavelength range from 450 to 700 nm and suitable for labeling proteins and amino-modified DNA. Journal of Chromatography B: Analytical Technologies in the Biomedical and Life Sciences, 793 (1), 83–92. Yuan, L., Lin, W., Yang, Y., and Chen, H. (2012) A unique class of near-infrared functional fluorescent dyes with carboxylic-acid-modulated fluorescence on/off switching: rational design, synthesis, optical properties, theoretical calculations, and applications for fluorescence imaging in living animals. Journal of the American Chemical Society, 134 (2), 1200–1211. Jacquemin, D. (2011) New cyanine dyes or not? Theoretical insights for model chains. The Journal of Physical Chemistry. A, 115 (11), 2442–2445. Ohira, S., Hales, J.M., Thorley, K.J., Anderson, H.L., Perry, J.W., and Bredas, J.-L. (2009) A new class of cyanine-like dyes with large bond-length alternation. Journal of the American Chemical Society, 131 (17), 6099–6101. Thorley, K.J., Hales, J.M., Anderson, H. L., and Perry, J.W. (2008) Porphyrin dimer carbocations with strong near infrared absorption and third-order optical nonlinearity. Angewandte Chemie, International Edition, 47 (37), 7095–7098. Zhao, J., Ji, S., Chen, Y., Guo, H., and Yang, P. (2012) Excited state intramolecular proton transfer (ESIPT): from principal photophysics to the development of new chromophores and applications in fluorescent molecular probes and luminescent materials. Physical Chemistry Chemical Physics, 14 (25), 8803–8817. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 238 45 Ma, J., Zhao, J., Yang, P., Huang, D., 46 47 48 49 50 51 52 53 54 55 56 Zhang, C., and Li, Q. (2012) New excited state intramolecular proton transfer (ESIPT) dyes based on naphthalimide and observation of long-lived triplet excited states. Chemical Communications, 48 (78), 9720–9722. Xie, L., Chen, Y., Wu, W., Guo, H., Zhao, J., and Yu, X. (2012) Fluorescent coumarin derivatives with large Stokes shift, dual emission and solid state luminescent properties: an experimental and theoretical study. Dyes and Pigments, 92 (3), 1361–1369. Lakowicz, J.R. (2006) Principles of Fluorescence Spectroscopy, 3rd edn, Springer. Tan, W.H., Wang, K.M., and Drake, T.J. (2004) Molecular beacons. Current Opinion in Chemical Biology, 8 (5), 547–553. Tan, L., Li, Y., Drake, T.J., Moroz, L., Wang, K.M., Li, J., Munteanu, A., Yang, C.Y.J., Martinez, K., and Tan, W.H. (2005) Molecular beacons for bioanalytical applications. Analyst, 130 (7), 1002–1005. Wu, P.G. and Brand, L. (1994) Resonance energy transfer: methods and applications. Analytical Biochemistry, 218 (1), 1–13. Buschmann, V., Weston, K.D., and Sauer, M. (2003) Spectroscopic study and evaluation of red-absorbing fluorescent dyes. Bioconjugate Chemistry, 14 (1), 195– 204. Escobedo, J.O., Rusin, O., Lim, S., and Strongin, R.M. (2010) NIR dyes for bioimaging applications. Current Opinion in Chemical Biology, 14 (1), 64–70. Carman, C.V. (2012) Overview: imaging in the study of integrins. Methods in Molecular Biology (Clifton, N. J.), 757, 159– 189. Loura, L.M.S. and Prieto, M. (2011) FRET in membrane biophysics: an overview. Frontiers in Physiology, 2, 82. Hayward, R.D., Goguen, J.D., and Leong, J.M. (2010) No better time to FRET: shedding light on host pathogen interactions. Journal of Biology (London), 9, 12. Gambin, Y. and Deniz, A.A. (2010) Multicolor single-molecule FRET to 57 58 59 60 61 62 63 64 65 66 explore protein folding and binding. Molecular Biosystems, 6 (9), 1540–1547. Vaddiraju, S., Burgess, D.J., Tomazos, I., Jain, F.C., and Papadimitrakopoulos, F. (2010) Technologies for continuous glucose monitoring: current problems and future promises. Journal of Diabetes Science and Technology, 4 (6), 1540–1562. Sapsford, K.E., Sun, S., Francis, J., Sharma, S., Kostov, Y., and Rasooly, A. (2008) A fluorescence detection platform using spatial electroluminescent excitation for measuring botulinum neurotoxin a activity. Biosensors & Bioelectronics, 24 (4), 618–625. Ahmad, A.I. and Ghasemi, J.B. (2007) New FRET primers for quantitative realtime PCR. Analytical and Bioanalytical Chemistry, 387 (8), 2737–2743. Juskowiak, B. (2011) Nucleic acid-based fluorescent probes and their analytical potential. Analytical and Bioanalytical Chemistry, 399 (9), 3157–3176. Waggoner, A. (2006) Fluorescent labels for proteomics and genomics. Current Opinion in Chemical Biology, 10 (1), 62–66. Li, H., Luo, Y., and Sun, X. (2011) Fluorescence resonance energy transfer dye-labeled probe for fluorescenceenhanced DNA detection: an effective strategy to greatly improve discrimination ability toward single-base mismatch. Biosensors & Bioelectronics, 27 (1), 167–171. Li, Y., Zhou, X., and Ye, D. (2008) Molecular beacons: an optimal multifunctional biological probe. Biochemical and Biophysical Research Communications, 373 (4), 457–461. Lewis, F.D., Zhang, L.G., and Zuo, X.B. (2005) Orientation control of fluorescence resonance energy transfer using DNA as a helical scaffold. Journal of the American Chemical Society, 127 (28), 10002–10003. Rindermann, J.J., Akhtman, Y., Richardson, J., Brown, T., and Lagoudakis, P.G. (2011) Gauging the flexibility of fluorescent markers for the interpretation of fluorescence resonance energy transfer. Journal of the American Chemical Society, 133 (2), 279–285. Sindbert, S., Kalinin, S., Hien, N., Kienzler, A., Clima, L., Bannwarth, W., j239 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 67 68 69 70 71 72 73 74 75 Appel, B., Mueller, S., and Seidel, C.A.M. (2011) Accurate distance determination of nucleic acids via Forster resonance energy transfer: implications of dye linker length and rigidity. Journal of the American Chemical Society, 133 (8), 2463–2480. Hassibi, A., Vikalo, H., Riechmann, J.L., and Hassibi, B. (2012) FRET-based realtime DNA microarrays. Methods in Molecular Biology (Clifton, N. J.), 815, 147–159. Peng, X., Chen, H., Draney, D.R., Volcheck, W., Schutz-Geschwender, A., and Olive, D.M. (2009) A nonfluorescent, broad-range quencher dye for Forster resonance energy transfer assays. Analytical Biochemistry, 388 (2), 220–228. Johnston, A.P.R. and Caruso, F. (2005) A molecular beacon approach to measuring the DNA permeability of thin films. Journal of the American Chemical Society, 127 (28), 10014–10015. Liu, J.W. and Lu, Y. (2003) Improving fluorescent DNAzyme biosensors by combining inter- and intramolecular quenchers. Analytical Chemistry, 75 (23), 6666–6672. Medintz, I.L., Goldman, E.R., Lassman, M.E., and Mauro, J.M. (2003) A fluorescence resonance energy transfer sensor based on maltose binding protein. Bioconjugate Chemistry, 14 (5), 909–918. Goldberg, J.M., Batjargal, S., and Petersson, E.J. (2010) Thioamides as fluorescence quenching probes: minimalist chromophores to monitor protein dynamics. Journal of the American Chemical Society, 132 (42), 14718–14720. Zhang, Z., Lu, X., Fan, Q., Hu, W., and Huang, W. (2011) Conjugated polyelectrolyte brushes with extremely high charge density for improved energy transfer and fluorescence quenching applications. Polymer Chemistry, 2 (10), 2369–2377. Loving, G.S., Sainlos, M., and Imperiali, B. (2010) Monitoring protein interactions and dynamics with solvatochromic fluorophores. Trends in Biotechnology, 28 (2), 73–83. Wagner, B.D. (2009) The use of coumarins as environmentally-sensitive fluorescent probes of heterogeneous 76 77 78 79 80 81 82 83 84 inclusion systems. Molecules, 14 (1), 210–237. Wang, R., Yu, C., Yu, F., and Chen, L. (2010) Molecular fluorescent probes for monitoring pH changes in living cells. Trends in Analytical Chemistry, 29 (9), 1004–1013. Wang, S., Li, N., Pan, W., and Tang, B. (2012) Advances in functional fluorescent and luminescent probes for imaging intracellular small-molecule reactive species. Trends in Analytical Chemistry, 39, 3–37. Bizzarri, R., Serresi, M., Luin, S., and Beltram, F. (2009) Green fluorescent protein based pH indicators for in vivo use: a review. Analytical and Bioanalytical Chemistry, 393 (4), 1107–1122. Dennis, A.M., Rhee, W.J., Sotto, D., Dublin, S.N., and Bao, G. (2012) Quantum dot-fluorescent protein FRET probes for sensing intracellular pH. ACS Nano, 6 (4), 2917–2924. Paradiso, A.M., Tsien, R.Y., and Machen, T.E. (1984) Naþ–Hþ exchange in gastric glands as measured with a cytoplasmictrapped, fluorescent pH indicator. Proceedings of the National Academy of Sciences of the United States of America, 81 (23), 7436–7440. Nakata, E., Nazumi, Y., Yukimachi, Y., Uto, Y., Maezawa, H., Hashimoto, T., Okamoto, Y., and Hori, H. (2011) Synthesis and photophysical properties of new SNARF derivatives as dual emission pH sensors. Bioorganic & Medicinal Chemistry Letters, 21 (6), 1663–1666. Vendrell, M., Zhai, D., Er, J.C., and Chang, Y.-T. (2012) Combinatorial strategies in fluorescent probe development. Chemical Reviews, 112 (8), 4391–4420. Chang, Q., Sipior, J., Lakowicz, J.R., and Rao, G. (1995) A lifetime-based fluorescence resonance energy-transfer sensor for ammonia. Analytical Biochemistry, 232 (1), 92–97. Lakowicz, J.R., Szmacinski, H., and Karakelle, M. (1993) Optical sensing of pH and PCO2 using phase-modulation fluorometry and resonance energytransfer. Analytica Chimica Acta, 272 (2), 179–186. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 240 85 Lippert, A.R., De Bittner, G.C.V., and 86 87 88 89 90 91 92 Chang, C.J. (2011) Boronate oxidation as a bioorthogonal reaction approach for studying the chemistry of hydrogen peroxide in living systems. Accounts of Chemical Research, 44 (9), 793–804. Albers, A.E., Okreglak, V.S., and Chang, C.J. (2006) A FRET-based approach to ratiometric fluorescence detection of hydrogen peroxide. Journal of the American Chemical Society, 128 (30), 9640–9641. Kurishita, Y., Kohira, T., Ojida, A., and Hamachi, I. (2010) Rational design of FRET-based ratiometric chemosensors for in vitro and in cell fluorescence analyses of nucleoside polyphosphates. Journal of the American Chemical Society, 132 (38), 13290–13299. Hu, J., Zhang, X., Wang, D., Hu, X., Liu, T., Zhang, G., and Liu, S. (2011) Ultrasensitive ratiometric fluorescent pH and temperature probes constructed from dye-labeled thermoresponsive double hydrophilic block copolymers. Journal of Materials Chemistry, 21 (47), 19030–19038. Childress, E.S., Roberts, C.A., Sherwood, D.Y., LeGuyader, C.L.M., and Harbron, E.J. (2012) Ratiometric fluorescence detection of mercury ions in water by conjugated polymer nanoparticles. Analytical Chemistry, 84 (3), 1235–1239. Mahato, P., Saha, S., Suresh, E., Di Liddo, R., Parnigotto, P.P., Conconi, M.T., Kesharwani, M.K., Ganguly, B., and Das, A. (2012) Ratiometric detection of Cr3þ and Hg2þ by a naphthalimide-rhodamine based fluorescent probe. Inorganic Chemistry, 51 (3), 1769–1777. Cosa, G., Focsaneanu, K.S., McLean, J.R. N., McNamee, J.P., and Scaiano, J.C. (2001) Photophysical properties of fluorescent DNA-dyes bound to singleand double-stranded DNA in aqueous buffered solution. Photochemistry and Photobiology, 73 (6), 585–599. Dragan, A.I., Pavlovic, R., McGivney, J.B., Casas-Finet, J.R., Bishop, E.S., Strouse, R. J., Schenerman, M.A., and Geddes, C.D. (2012) SYBR Green I: fluorescence properties and interaction with DNA. Journal of Fluorescence, 22 (4), 1189–1199. 93 Singh, M.P., Jennings, T.L., and Strouse, 94 95 96 97 98 99 100 101 G.F. (2009) Tracking spatial disorder in an optical ruler by time-resolved NSET. The Journal of Physical Chemistry. B, 113 (2), 552–558. Lopez, S.G., Ruedas-Rama, M.J., Casares, S., Alvarez-Pez, J.M., and Orte, A. (2012) Bulk and single-molecule fluorescence studies of the saturation of the DNA double helix using YOYO-3 intercalator dye. The Journal of Physical Chemistry. B, 116 (38), 11561–11569. Takatsu, K., Yokomaku, T., Kurata, S., and Kanagawa, T. (2004) A FRET-based analysis of SNPs without fluorescent probes. Nucleic Acids Research, 32 (19), e156. Halpern, M.D. and Ballantyne, J. (2009) A single nucleotide polymorphism melt curve assay employing an intercalating dye probe fluorescence resonance energy transfer for forensic analysis. Analytical Biochemistry, 391 (1), 1–10. Halpern, M.D. and Ballantyne, J. (2011) An STR melt curve genotyping assay for forensic analysis employing an intercalating dye probe FRET. Journal of Forensic Sciences, 56 (1), 36–45. Das, D.K., Das, A.K., Mondal, T., Mandal, A.K., and Bhattacharyya, K. (2010) Ultrafast FRET in ionic liquid-P123 mixed micelles: region and counterion dependence. The Journal of Physical Chemistry. B, 114 (41), 13159–13166. Chen, J., Zhang, P., Fang, G., Yi, P., Zeng, F., and Wu, S. (2012) Design and synthesis of FRET-mediated multicolor and photoswitchable fluorescent polymer nanoparticles with tunable emission properties. The Journal of Physical Chemistry. B, 116 (14), 4354–4362. Chen, J., Zhang, P., and Yi, P. (2010) Preparation of fluorescence tunable polymer nanoparticles by one-step miniemulsion. Journal of Macromolecular Science, Part A: Pure and Applied Chemistry, 47 (11), 1135–1141. Ha, S.-W., Camalier, C.E., Beck, Jr., G.R., and Lee, J.-K. (2009) New method to prepare very stable and biocompatible fluorescent silica nanoparticles. Chemical Communications, (20), 2881–2883. j241 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 102 Wagh, A., Qian, S.Y., and Law, B. (2012) 103 104 105 106 107 108 109 110 111 Development of biocompatible polymeric nanoparticles for in vivo NIR and FRET imaging. Bioconjugate Chemistry, 23 (5), 981–992. Ueno, Y., Jose, J., Loudet, A., PerezBolivar, C., Anzenbacher, Jr., P., and Burgess, K. (2011) Encapsulated energytransfer cassettes with extremely well resolved fluorescent outputs. Journal of the American Chemical Society, 133 (1), 51–55. Wu, C.F., Szymanski, C., and McNeill, J. (2006) Preparation and encapsulation of highly fluorescent conjugated polymer nanoparticles. Langmuir, 22 (7), 2956–2960. Wang, L., Zhao, W., and Tan, W. (2008) Bioconjugated silica nanoparticles: development and applications. Nano Research, 1 (2), 99–115. Jiang, S., Gnanasammandhan, M.K., and Zhang, Y. (2010) Optical imaging-guided cancer therapy with fluorescent nanoparticles. Journal of the Royal Society Interface, 7 (42), 3–18. Valanne, A., Lindroos, H., Lovgren, T., and Soukka, T. (2005) A novel homogeneous assay format utilising proximity dependent fluorescence energy transfer between particulate labels. Analytica Chimica Acta, 539 (1–2), 251–256. Roberts, D.V., Wittmershaus, B.P., Zhang, Y.Z., Swan, S., and Klinosky, M.P. (1998) Efficient excitation energy transfer among multiple dyes in polystyrene microspheres. Journal of Luminescence, 79 (4), 225–231. Wu, W.-B., Wang, M.-L., Sun, Y.-M., Huang, W., Cui, Y.-P., and Xu, C.-X. (2008) Color-tuned FRET polystyrene microspheres by single wavelength excitation. Optical Materials, 30 (12), 1803–1809. Wang, L. and Tan, W.H. (2006) Multicolor FRET silica nanoparticles by single wavelength excitation. Nano Letters, 6 (1), 84–88. Wang, L., Zhao, W., O’Donoghue, M.B., and Tan, W. (2007) Fluorescent nanoparticles for multiplexed bacteria monitoring. Bioconjugate Chemistry, 18 (2), 297–301. 112 Lessard-Viger, M., Rioux, M., Rainville, L., 113 114 115 116 117 118 119 120 and Boudreau, D. (2009) FRET enhancement in multilayer core–shell nanoparticles. Nano Letters, 9 (8), 3066–3071. Yang, J., Zhang, F., Chen, Y., Qian, S., Hu, P., Li, W., Deng, Y., Fang, Y., Han, L., Luqman, M., and Zhao, D. (2011) Core– shell Ag@SiO2@mSiO(2) mesoporous nanocarriers for metal-enhanced fluorescence. Chemical Communications, 47 (42), 11618–11620. Ow, H., Larson, D.R., Srivastava, M., Baird, B.A., Webb, W.W., and Wiesner, U. (2005) Bright and stable core–shell fluorescent silica nanoparticles. Nano Letters, 5 (1), 113–117. Cohen, B., Martin, C., Iyer, S.K., Wiesner, U., and Douhal, A. (2012) Single dye molecule behavior in fluorescent core– shell silica nanoparticles. Chemistry of Materials, 24 (2), 361–372. Seelig, J., Leslie, K., Renn, A., Kuhn, S., Jacobsen, V., van de Corput, M., Wyman, C., and Sandoghdar, V. (2007) Nanoparticle-induced fluorescence lifetime modification as nanoscopic ruler: demonstration at the single molecule level. Nano Letters, 7 (3), 685–689. Szollosi, J., Damjanovich, S., and Matyus, L. (1998) Application of fluorescence resonance energy transfer in the clinical laboratory: routine and research. Cytometry, 34 (4), 159–179. Horvath, G., Petras, M., Szentesi, G., Fabian, A., Park, J.W., Vereb, G., and Szollosi, J. (2005) Selecting the right fluorophores and flow cytometer for fluorescence resonance energy transfer measurements. Cytometry. Part A: the Journal of the International Society for Analytical Cytology, 65 (2), 148–157. Rao, K.V.N., Stevens, P.W., Hall, J.G., Lyamichev, V., Neri, B.P., and Kelso, D.M. (2003) Genotyping single nucleotide polymorphisms directly from genomic DNA by invasive cleavage reaction on microspheres. Nucleic Acids Research, 31 (11), e66. Nolan, J.P. and Mandy, F. (2006) Multiplexed and microparticle-based analyses: quantitative tools for the largescale analysis of biological systems. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 242 121 122 123 124 125 126 127 128 Cytometry. Part A: the Journal of the International Society for Analytical Cytology, 69 (5), 318–325. Tran, P.L., Gamboa, J.R., You, D.J., and Yoon, J.-Y. (2010) FRET detection of Octamer-4 on a protein nanoarray made by size-dependent self-assembly. Analytical and Bioanalytical Chemistry, 398 (2), 759–768. Trinchi, A. and Muster, T.H. (2007) A review of surface functionalized amine terminated dendrimers for application in biological and molecular sensing. Supramolecular Chemistry, 19 (7), 431–445. Menjoge, A.R., Kannan, R.M., and Tomalia, D.A. (2010) Dendrimer-based drug and imaging conjugates: design considerations for nanomedical applications. Drug Discovery Today, 15 (5– 6), 171–185. D’Ambruoso, G.D. and McGrath, D.V. (2008) Energy harvesting in synthetic dendrimer materials, in Photoresponsive Polymers Ii (eds S.R. Marder and K.S. Lee), Springer Berlin Heidelberg, pp. 87–147. Zaragoza-Galan, G., Fowler, M.A., Duhamel, J., Rein, R., Solladie, N., and Rivera, E. (2012) Synthesis and characterization of novel pyrenedendronized porphyrins exhibiting efficient fluorescence resonance energy transfer: optical and photophysical properties. Langmuir, 28 (30), 11195–11205. Szent-Gyorgyi, C., Schmidt, B.F., Fitzpatrick, J.A.J., and Bruchez, M.P. (2010) Fluorogenic dendrons with multiple donor chromophores as bright genetically targeted and activated probes. Journal of the American Chemical Society, 132 (32), 11103–11109. Rashatasakhon, P., Vongnam, K., Siripornnoppakhun, W., Vilaivan, T., and Sukwattanasinitt, M. (2012) FRET detection of DNA sequence via electrostatic interaction of polycationic phenyleneethynylene dendrimer with DNA/PNA hybrid. Talanta, 88, 593–598. Davis, B.W., Niamnont, N., Dillon, R., Bardeen, C.J., Sukwattanasinitt, M., and Cheng, Q. (2011) FRET detection of proteins using fluorescently doped 129 130 131 132 133 134 135 136 137 electrospun nanofibers and pattern recognition. Langmuir, 27 (10), 6401–6408. Myc, A., Majoros, I.J., Thomas, T.P., and Baker, Jr., J.R. (2007) Dendrimer-based targeted delivery of an apoptotic sensor in cancer cells. Biomacromolecules, 8 (1), 13–18. Alvarez, A., Salinas-Castillo, A., CostaFernandez, J.M., Pereiro, R., and SanzMedel, A. (2011) Fluorescent conjugated polymers for chemical and biochemical sensing. Trends in Analytical Chemistry, 30 (9), 1513–1525. Wohlgenannt, M., Graupner, W., Wenzl, F.P., Tasch, S., List, E.J.W., Leising, G., Graupner, M., Hermetter, A., Rohr, U., Schlichting, P., Geerts, Y., Scherf, U., and Mullen, K. (1998) Photophysics of excitation energy transfer in highly fluorescent polymers. Chemical Physics, 227 (1–2), 99–109. Liu, X., Fan, Q., and Huang, W. (2011) DNA biosensors based on water-soluble conjugated polymers. Biosensors & Bioelectronics, 26 (5), 2154–2164. Wang, B., Yang, Q., Liu, L., and Wang, S. (2011) Direct energy transfer from conjugated polymer to DNA intercalated dye: label-free fluorescent DNA detection. Colloids and Surfaces B: Biointerfaces, 85 (1), 8–11. Srinivas, A.R.G., Peng, H., Barker, D., and Travas-Sejdic, J. (2012) Switch on or switch off: an optical DNA sensor based on poly(p-phenylenevinylene) grafted magnetic beads. Biosensors & Bioelectronics, 35 (1), 498–502. Jiang, G., Susha, A.S., Lutich, A.A., Stefani, F.D., Feldmann, J., and Rogach, A.L. (2009) Cascaded FRET in conjugated polymer/quantum dot/dye-labeled DNA complexes for DNA hybridization detection. ACS Nano, 3 (12), 4127–4131. Feng, F., Liu, L., and Wang, S. (2010) Fluorescent conjugated polymer-based FRET technique for detection of DNA methylation of cancer cells. Nature Protocols, 5 (7), 1255–1264. Cheng, Y., Du, Q., Wang, L., Jia, H., and Li, Z. (2012) Fluorescently cationic conjugated polymer as an indicator of ligase chain reaction for sensitive and j243 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 138 139 140 141 142 143 144 145 homogeneous detection of single nucleotide polymorphism. Analytical Chemistry, 84 (8), 3739–3744. Sun, P., Lu, X., Fan, Q., Zhang, Z., Song, W., Li, B., Huang, L., Peng, J., and Huang, W. (2011) Water-soluble iridium(III)containing conjugated polyelectrolytes with weakened energy transfer properties for multicolor protein sensing applications. Macromolecules, 44 (22), 8763–8770. Pu, K.-Y., Luo, Z., Li, K., Xie, J., and Liu, B. (2011) Energy transfer between conjugated-oligoelectrolyte-substituted POSS and gold nanocluster for multicolor intracellular detection of mercury ion. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 115 (26), 13069–13075. Wu, C., Bull, B., Szymanski, C., Christensen, K., and McNeill, J. (2008) Multicolor conjugated polymer dots for biological fluorescence imaging. ACS Nano, 2 (11), 2415–2423. Wu, C., Schneider, T., Zeigler, M., Yu, J., Schiro, P.G., Burnham, D.R., McNeill, J. D., and Chiu, D.T. (2010) Bioconjugation of ultrabright semiconducting polymer dots for specific cellular targeting. Journal of the American Chemical Society, 132 (43), 15410–15417. Wu, C., Szymanski, C., Cain, Z., and McNeill, J. (2007) Conjugated polymer dots for multiphoton fluorescence imaging. Journal of the American Chemical Society, 129 (43), 12904–12905. Chan, Y.-H., Gallina, M.E., Zhang, X., Wu, I.C., Jin, Y., Sun, W., and Chiu, D.T. (2012) Reversible photoswitching of spiropyran-conjugated semiconducting polymer dots. Analytical Chemistry, 84 (21), 9431–9438. Chan, Y.-H., Wu, C., Ye, F., Jin, Y., Smith, P.B., and Chiu, D.T. (2011) Development of ultrabright semiconducting polymer dots for ratiometric pH sensing. Analytical Chemistry, 83 (4), 1448–1455. Clafton, S.N., Beattie, D.A., MierczynskaVasilev, A., Acres, R.G., Morgan, A.C., and Kee, T.W. (2010) Chemical defects in the highly fluorescent conjugated polymer dots. Langmuir, 26 (23), 17785–17789. 146 Alfimov, M.V., Fedorova, O.A., and 147 148 149 150 151 152 153 154 155 Gromov, S.P. (2003) Photoswitchable molecular receptors. Journal of Photochemistry and Photobiology A: Chemistry, 158 (2–3), 183–198. Pardo, R., Zayat, M., and Levy, D. (2011) Photochromic organic-inorganic hybrid materials. Chemical Society Reviews, 40 (2), 672–687. Delaire, J.A. and Nakatani, K. (2000) Linear and nonlinear optical properties of photochromic molecules and materials. Chemical Reviews, 100 (5), 1817–1845. Cusido, J., Deniz, E., and Raymo, F.M. (2009) Fluorescent switches based on photochromic compounds. European Journal of Organic Chemistry, 2009 (13), 2031–2045. Yun, C., You, J., Kim, J., Huh, J., and Kim, E. (2009) Photochromic fluorescence switching from diarylethenes and its applications. Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 10 (3), 111–129. Berkovic, G., Krongauz, V., and Weiss, V. (2000) Spiropyrans and spirooxazines for memories and switches. Chemical Reviews, 100 (5), 1741–1753. Giordano, L., Jovin, T.M., Irie, M., and Jares-Erijman, E.A. (2002) Diheteroarylethenes as thermally stable photoswitchable acceptors in photochromic fluorescence resonance energy transfer (pcFRET). Journal of the American Chemical Society, 124 (25), 7481– 7489. Piao, X., Zou, Y., Wu, J., Li, C., and Yi, T. (2009) Multiresponsive switchable diarylethene and its application in bioimaging. Organic Letters, 11 (17), 3818–3821. Yildiz, I., Deniz, E., and Raymo, F.M. (2009) Fluorescence modulation with photochromic switches in nanostructured constructs. Chemical Society Reviews, 38 (7), 1859–1867. Stepanenko, O.V., Stepanenko, O.V., Shcherbakova, D.M., Kuznetsova, I.M., Turoverov, K.K., and Verkhusha, V.V. (2011) Modern fluorescent proteins: from chromophore formation to novel intracellular applications. Biotechniques, 51 (5), 313–318. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 244 156 Yan, Y., Marriott, M.E., Petchprayoon, C., 157 158 159 160 161 162 163 and Marriott, G. (2011) Optical switch probes and optical lock-in detection (OLID) imaging microscopy: highcontrast fluorescence imaging within living systems. Biochemical Journal, 433, 411–422. Mao, S., Benninger, R.K.P., Yan, Y., Petchprayoon, C., Jackson, D., Easley, C.J., Piston, D.W., and Marriott, G. (2008) Optical lock-in detection of FRET using synthetic and genetically encoded optical switches. Biophysical Journal, 94 (11), 4515–4524. Marriott, G., Mao, S., Sakata, T., Ran, J., Jackson, D.K., Petchprayoon, C., Gomez, T.J., Warp, E., Tulyathan, O., Aaron, H.L., Isacoff, E.Y., and Yan, Y. (2008) Optical lock-in detection imaging microscopy for contrast-enhanced imaging in living cells. Proceedings of the National Academy of Sciences of the United States of America, 105 (46), 17789–17794. Petchprayoon, C., Yan, Y., Mao, S., and Marriott, G. (2011) Rational design, synthesis, and characterization of highly fluorescent optical switches for highcontrast optical lock-in detection (OLID) imaging microscopy in living cells. Bioorganic & Medicinal Chemistry, 19 (3), 1030–1040. Wang, Y., Hong, C.-Y., and Pan, C.-Y. (2012) Spiropyran-based hyperbranched star copolymer: synthesis, phototropy, FRET, and bioapplication. Biomacromolecules, 13 (8), 2585–2593. Kruttwig, K., Yankelevich, D.R., Brueggemann, C., Tu, C., L’Etoile, N., Knoesen, A., and Louie, A.Y. (2012) Reversible low-light induced photoswitching of crowned spiropyranDO3A complexed with Gadolinium(III) ions. Molecules, 17 (6), 6605–6624. Medintz, I.L., Trammell, S.A., Mattoussi, H., and Mauro, J.M. (2004) Reversible modulation of quantum dot photoluminescence using a proteinbound photochromic fluorescence resonance energy transfer acceptor. Journal of the American Chemical Society, 126 (1), 30–31. Kim, S., Yoon, S.-J., and Park, S.Y. (2012) Highly fluorescent chameleon 164 165 166 167 168 169 170 171 172 nanoparticles and polymer films: multicomponent organic systems that combine FRET and photochromic switching. Journal of the American Chemical Society, 134 (29), 12091–12097. Zhu, L., Wu, W., Zhu, M.-Q., Han, J.J., Hurst, J.K., and Li, A.D.Q. (2007) Reversibly photoswitchable dual-color fluorescent nanoparticles as new tools for live-cell imaging. Journal of the American Chemical Society, 129 (12), 3524–3526. Diaz, S.A., Giordano, L., Jovin, T.M., and Jares-Erijman, E.A. (2012) Modulation of a photoswitchable dual-color quantum dot containing a photochromic FRET acceptor and an internal standard. Nano Letters, 12 (7), 3537–3544. Diaz, S.A., Menendez, G.O., Etchehon, M.H., Giordano, L., Jovin, T.M., and JaresErijman, E.A. (2011) Photoswitchable water-soluble quantum dots: pcFRET based on amphiphilic photochromic polymer coating. ACS Nano, 5 (4), 2795–2805. Kim, Y., Jung, H.-y., Choe, Y.H., Lee, C., Ko, S.-K., Koun, S., Choi, Y., Chung, B.H., Park, B.C., Huh, T.-L., Shin, I., and Kim, E. (2012) High-contrast reversible fluorescence photoswitching of dyecrosslinked dendritic nanoclusters in living vertebrates. Angewandte Chemie, International Edition, 51 (12), 2878–2882. Newman, R.H., Fosbrink, M.D., and Zhang, J. (2011) Genetically encodable fluorescent biosensors for tracking signaling dynamics in living cells. Chemical Reviews, 111 (5), 3614–3666. Subach, F.V. and Verkhusha, V.V. (2012) Chromophore transformations in red fluorescent proteins. Chemical Reviews, 112 (7), 4308–4327. Kochmann, S., Hirsch, T., and Wolfbeis, O.S. (2012) Graphenes in chemical sensors and biosensors. Trends in Analytical Chemistry, 39, 87–113. Dong, H., Gao, W., Yan, F., Ji, H., and Ju, H. (2010) Fluorescence resonance energy transfer between quantum dots and graphene oxide for sensing biomolecules. Analytical Chemistry, 82 (13), 5511–5517. Chang, H., Tang, L., Wang, Y., Jiang, J., and Li, J. (2010) Graphene fluorescence resonance energy transfer aptasensor for j245 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 173 174 175 176 177 178 179 180 the thrombin detection. Analytical Chemistry, 82 (6), 2341–2346. Wu, S., Duan, N., Ma, X., Xia, Y., Wang, H., Wang, Z., and Zhang, Q. (2012) Multiplexed fluorescence resonance energy transfer aptasensor between upconversion nanoparticles and graphene oxide for the simultaneous determination of mycotoxins. Analytical Chemistry, 84 (14), 6263–6270. Cao, L., Cheng, L., Zhang, Z., Wang, Y., Zhang, X., Chen, H., Liu, B., Zhang, S., and Kong, J. (2012) Visual and highthroughput detection of cancer cells using a graphene oxide-based FRET aptasensing microfluidic chip. Lab on a Chip, 12 (22), 4864–4869. Zhao, X.-H., Kong, R.-M., Zhang, X.-B., Meng, H.-M., Liu, W.-N., Tan, W., Shen, G.-L., and Yu, R.-Q. (2011) GrapheneDNAzyme based biosensor for amplified fluorescence “turn-on” detection of Pb2þ with a high selectivity. Analytical Chemistry, 83 (13), 5062–5066. Chien, C.-T., Li, S.-S., Lai, W.-J., Yeh, Y.-C., Chen, H.-A., Chen, I.S., Chen, L.-C., Chen, K.-H., Nemoto, T., Isoda, S., Chen, M., Fujita, T., Eda, G., Yamaguchi, H., Chhowalla, M., and Chen, C.-W. (2012) Tunable photoluminescence from graphene oxide. Angewandte Chemie, International Edition, 51 (27), 6662–6666. Liu, F., Choi, J.Y., and Seo, T.S. (2010) Graphene oxide arrays for detecting specific DNA hybridization by fluorescence resonance energy transfer. Biosensors & Bioelectronics, 25 (10), 2361–2365. Shi, Y., Wu, J., Sun, Y., Zhang, Y., Wen, Z., Dai, H., Wang, H., and Li, Z. (2012) A graphene oxide based biosensor for microcystins detection by fluorescence resonance energy transfer. Biosensors & Bioelectronics, 38 (1), 31–36. Liu, R., Wu, D., Feng, X., and Muellen, K. (2011) Bottom-up fabrication of photoluminescent graphene quantum dots with uniform morphology. Journal of the American Chemical Society, 133 (39), 15221–15223. Zhuo, S., Shao, M., and Lee, S.-T. (2012) Upconversion and downconversion fluorescent graphene quantum dots: 181 182 183 184 185 186 187 188 189 190 ultrasonic preparation and photocatalysis. ACS Nano, 6 (2), 1059–1064. Fan, L., Hu, Y., Wang, X., Zhang, L., Li, F., Han, D., Li, Z., Zhang, Q., Wang, Z., and Niu, L. (2012) Fluorescence resonance energy transfer quenching at the surface of graphene quantum dots for ultrasensitive detection of TNT. Talanta, 101, 192–197. Welsher, K., Liu, Z., Sherlock, S.P., Robinson, J.T., Chen, Z., Daranciang, D., and Dai, H. (2009) A route to brightly fluorescent carbon nanotubes for nearinfrared imaging in mice. Nature Nanotechnology, 4 (11), 773–780. Harris, P.J.F. (2011) Carbon Nantube Science: Synthesis, Properties and Applications, 2nd edn, Cambridge University Press. Shafran, E., Mangum, B.D., and Gerton, J. M. (2010) Energy transfer from an individual quantum dot to a carbon nanotube. Nano Letters, 10 (10), 4049–4054. Chiu, C.F., Dementev, N., and Borguet, E. (2011) Fluorescence quenching of dyes covalently attached to single-walled carbon nanotubes. The Journal of Physical Chemistry. A, 115 (34), 9579–9584. Ignatova, T., Najafov, H., Ryasnyanskiy, A., Biaggio, I., Zheng, M., and Rotkin, S. V. (2011) Significant FRET between SWNT/DNA and rare earth ions: a signature of their spatial correlations. ACS Nano, 5 (7), 6052–6059. Tian, J., Zhao, H., Liu, M., Chen, Y., and Quan, X. (2012) Detection of influenza a virus based on fluorescence resonance energy transfer from quantum dots to carbon nanotubes. Analytica Chimica Acta, 723, 83–87. Baker, S.N. and Baker, G.A. (2010) Luminescent carbon nanodots: emergent nanolights. Angewandte Chemie, International Edition, 49 (38), 6726–6744. Liu, J.-H., Yang, S.-T., Chen, X.-X., and Wang, H. (2012) Fluorescent carbon dots and nanodiamonds for biological imaging: preparation, application, pharmacokinetics and toxicity. Current Drug Metabolism, 13 (8), 1046–1056. Schrand, A.M., Hens, S.A.C., and Shenderova, O.A. (2009) Nanodiamond Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 246 191 192 193 194 195 196 197 198 199 particles: properties and perspectives for bioapplications. Critical Reviews in Solid State and Materials Sciences, 34 (1–2), 18–74. Mochalin, V.N., Shenderova, O., Ho, D., and Gogotsi, Y. (2012) The properties and applications of nanodiamonds. Nature Nanotechnology, 7 (1), 11–23. Wang, Y., Bao, L., Liu, Z., and Pang, D.-W. (2011) Aptamer biosensor based on fluorescence resonance energy transfer from upconverting phosphors to carbon nanoparticles for thrombin detection in human plasma. Analytical Chemistry, 83 (21), 8130–8137. Wang, Y., Shen, P., Li, C., Wang, Y., and Liu, Z. (2012) Upconversion fluorescence resonance energy transfer based biosensor for ultrasensitive detection of matrix metalloproteinase-2 in blood. Analytical Chemistry, 84 (3), 1466–1473. Tisler, J., Reuter, R., Laemmle, A., Jelezko, F., Balasubramanian, G., Hemmer, P.R., Reinhard, F., and Wrachtrup, J. (2011) Highly efficient FRET from a single nitrogen-vacancy center in nanodiamonds to a single organic molecule. ACS Nano, 5 (10), 7893–7898. Chen, Y.-Y., Shu, H., Kuo, Y., Tzeng, Y.-K., and Chang, H.-C. (2011) Measuring Forster resonance energy transfer between fluorescent nanodiamonds and near-infrared dyes by acceptor photobleaching. Diamond and Related Materials, 20 (5–6), 803–807. Say, J.M., Bradac, C., Gaebel, T., Rabeau, J.R., and Brown, L.J. (2012) Processing 15nm nanodiamonds containing nitrogenvacancy centres for single-molecule FRET. Australian Journal of Chemistry, 65 (5), 496–503. Yengo, C.M. and Berger, C.L. (2010) Fluorescence anisotropy and resonance energy transfer: powerful tools for measuring real time protein dynamics in a physiological environment. Current Opinion in Pharmacology, 10 (6), 731–737. Demchenko, A.P. (1986) Ultraviolet Spectroscopy of Proteins, 1st edn, Springer. Kayser, V., Chennamsetty, N., Voynov, V., Helk, B., and Trout, B.L. (2011) Tryptophan–tryptophan energy transfer and classification of tryptophan residues 200 201 202 203 204 205 206 207 in proteins using a therapeutic monoclonal antibody as a model. Journal of Fluorescence, 21 (1), 275–288. Glasscock, J.M., Zhu, Y., Chowdhury, P., Tang, J., and Gai, F. (2008) Using an amino acid fluorescence resonance energy transfer pair to probe protein unfolding: application to the villin headpiece subdomain and the LysM domain. Biochemistry, 47 (42), 11070–11076. Visser, N.V., Westphal, A.H., van Hoek, A., van Mierlo, C.P.M., Visser, A.J.W.G., and van Amerongen, H. (2008) Tryptophan–tryptophan energy migration as a tool to follow apoflavodoxin folding. Biophysical Journal, 95 (5), 2462–2469. Jha, S.K., Dhar, D., Krishnamoorthy, G., and Udgaonkar, J.B. (2009) Continuous dissolution of structure during the unfolding of a small protein. Proceedings of the National Academy of Sciences of the United States of America, 106 (27), 11113–11118. Smirnova, I., Kasho, V., Sugihara, J., and Kaback, H.R. (2009) Probing of the rates of alternating access in LacY with Trp fluorescence. Proceedings of the National Academy of Sciences of the United States of America, 106 (51), 21561–21566. Lee, J.C., Langen, R., Hummel, P.A., Gray, H.B., and Winkler, J.R. (2004) Alpha-synuclein structures from fluorescence energy-transfer kinetics: implications for the role of the protein in Parkinson’s disease. Proceedings of the National Academy of Sciences of the United States of America, 101 (47), 16466–16471. Lovell, J.F., Chan, M.W., Qi, Q., Chen, J., and Zheng, G. (2011) Porphyrin FRET acceptors for apoptosis induction and monitoring. Journal of the American Chemical Society, 133 (46), 18580–18582. Ishida, Y., Shimada, T., Masui, D., Tachibana, H., Inoue, H., and Takagi, S. (2011) Efficient excited energy transfer reaction in clay/porphyrin complex toward an artificial light-harvesting system. Journal of the American Chemical Society, 133 (36), 14280–14286. Tucker, M.J., Oyola, R., and Gai, F. (2005) Conformational distribution of a 14residue peptide in solution: a fluorescence j247 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 208 209 210 211 212 213 214 215 resonance energy transfer study. The Journal of Physical Chemistry. B, 109 (10), 4788–4795. Kajihara, D., Abe, R., Iijima, I., Komiyama, C., Sisido, M., and Hohsaka, T. (2006) FRET analysis of protein conformational change through positionspecific incorporation of fluorescent amino acids. Nature Methods, 3 (11), 923–929. Connor, R.E. and Tirrell, D.A. (2007) Noncanonical amino acids in protein polymer design. Polymer Reviews, 47 (1), 9–28. Rogers, J.M.G., Lippert, L.G., and Gai, F. (2010) Non-natural amino acid fluorophores for one- and two-step fluorescence resonance energy transfer applications. Analytical Biochemistry, 399 (2), 182–189. Serrano, A.L., Troxler, T., Tucker, M.J., and Gai, F. (2010) Photophysics of a fluorescent non-natural amino acid: pcyanophenylalanine. Chemical Physics Letters, 487 (4–6), 303–306. Sarkar, S.S., Udgaonkar, J.B., and Krishnamoorthy, G. (2011) Reduced fluorescence lifetime heterogeneity of 5fluorotryptophan in comparison to tryptophan in proteins: implication for resonance energy transfer experiments. The Journal of Physical Chemistry. B, 115 (22), 7479–7486. Iijima, I. and Hohsaka, T. (2009) Positionspecific incorporation of fluorescent nonnatural amino acids into maltose-binding protein for detection of ligand binding by FRET and fluorescence quenching. ChemBioChem, 10 (6), 999–1006. Aprilakis, K.N., Taskent, H., and Raleighj, D.P. (2007) Use of the novel fluorescent amino acid rho-cyanophenylalanine offers a direct probe of hydrophobic core formation during the folding of the n-terminal domain of the ribosomal protein L9 and provides evidence for twostate folding. Biochemistry, 46 (43), 12308–12313. Taskent-Sezgin, H., Chung, J., Patsalo, V., Miyake-Stoner, S.J., Miller, A.M., Brewer, S.H., Mehl, R.A., Green, D.F., Raleigh, D. P., and Carrico, I. (2009) Interpretation of p-cyanophenylalanine fluorescence in proteins in terms of solvent exposure and 216 217 218 219 220 221 222 223 224 contribution of side-chain quenchers: a combined fluorescence, IR and molecular dynamics study. Biochemistry, 48 (38), 9040–9046. Miyake-Stoner, S.J., Miller, A.M., Hammill, J.T., Peeler, J.C., Hess, K.R., Mehl, R.A., and Brewer, S.H. (2009) Probing protein folding using sitespecifically encoded unnatural amino acids as FRET donors with tryptophan. Biochemistry, 48 (25), 5953–5962. Kuhn, S.M., Rubini, M., Mueller, M.A., and Skerra, A. (2011) Biosynthesis of a fluorescent protein with extreme pseudoStokes shift by introducing a genetically encoded non-natural amino acid outside the fluorophore. Journal of the American Chemical Society, 133 (11), 3708–3711. Shaner, N.C., Steinbach, P.A., and Tsien, R.Y. (2005) A guide to choosing fluorescent proteins. Nature Methods, 2 (12), 905–909. Giepmans, B.N.G., Adams, S.R., Ellisman, M.H., and Tsien, R.Y. (2006) The fluorescent toolbox for assessing protein location and function. Science, 312 (5771), 217–224. Palmer, A.E. and Tsien, R.Y. (2006) Measuring calcium signaling using genetically targetable fluorescent indicators. Nature Protocols, 1 (3), 1057–1065. Tsien, R.Y. (1998) The green fluorescent protein. Annual Review of Biochemistry, 67, 509–544. Shaner, N.C., Campbell, R.E., Steinbach, P.A., Giepmans, B.N.G., Palmer, A.E., and Tsien, R.Y. (2004) Improved monomeric red, orange and yellow fluorescent proteins derived from Discosoma sp. red fluorescent protein. Nature Biotechnology, 22 (12), 1567–1572. Shaner, N.C., Lin, M.Z., McKeown, M.R., Steinbach, P.A., Hazelwood, K.L., Davidson, M.W., and Tsien, R.Y. (2008) Improving the photostability of bright monomeric orange and red fluorescent proteins. Nature Methods, 5 (6), 545–551. Lam, A.J., St-Pierre, F., Gong, Y., Marshall, J.D., Cranfill, P.J., Baird, M.A., McKeown, M.R., Wiedenmann, J., Davidson, M.W., Schnitzer, M.J., Tsien, R. Y., and Lin, M.Z. (2012) Improving FRET Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 248 225 226 227 228 229 230 231 232 dynamic range with bright green and red fluorescent proteins. Nature Methods, 9 (10), 1005–1012. Kawano, H., Kogure, T., Abe, Y., Mizuno, H., and Miyawaki, A. (2008) Two-photon dual-color imaging using fluorescent proteins. Nature Methods, 5 (5), 373–374. Drobizhev, M., Makarov, N.S., Tillo, S.E., Hughes, T.E., and Rebane, A. (2011) Twophoton absorption properties of fluorescent proteins. Nature Methods, 8 (5), 393–399. Shemiakina, I.I., Ermakova, G.V., Cranfill, P.J., Baird, M.A., Evans, R.A., Souslova, E. A., Staroverov, D.B., Gorokhovatsky, A.Y., Putintseva, E.V., Gorodnicheva, T.V., Chepurnykh, T.V., Strukova, L., Lukyanov, S., Zaraisky, A.G., Davidson, M.W., Chudakov, D.M., and Shcherbo, D. (2012) A monomeric red fluorescent protein with low cytotoxicity. Nature Communications, 3, 1204. Andresen, M., Stiel, A.C., Foelling, J., Wenzel, D., Schoenle, A., Egner, A., Eggeling, C., Hell, S.W., and Jakobs, S. (2008) Photoswitchable fluorescent proteins enable monochromatic multilabel imaging and dual color fluorescence nanoscopy. Nature Biotechnology, 26 (9), 1035–1040. Piston, D.W. and Kremers, G.-J. (2007) Fluorescent protein FRET: the good, the bad and the ugly. Trends in Biochemical Sciences, 32 (9), 407–414. Boeneman, K., Delehanty, J.B., Susumu, K., Stewart, M.H., Deschamps, J.R., and Medintz, I.L. (2012) Quantum dots and fluorescent protein FRET-based biosensors. Advances in Experimental Medicine and Biology, 733, 63–74. Galban, J., Sanz-Vicente, I., Ortega, E., del Barrio, M., and de Marcos, S. (2012) Reagentless fluorescent biosensors based on proteins for continuous monitoring systems. Analytical and Bioanalytical Chemistry, 402 (10), 3039–3054. Hou, B.-H., Takanaga, H., Grossmann, G., Chen, L.-Q., Qu, X.-Q., Jones, A.M., Lalonde, S., Schweissgut, O., Wiechert, W., and Frommer, W.B. (2011) Optical sensors for monitoring dynamic changes of intracellular metabolite levels in 233 234 235 236 237 238 239 240 241 242 mammalian cells. Nature Protocols, 6 (11), 1818–1833. Wegner, S.V., Arslan, H., Sunbul, M., Yin, J., and He, C. (2010) Dynamic copper(I) imaging in mammalian cells with a genetically encoded fluorescent copper(I) sensor. Journal of the American Chemical Society, 132 (8), 2567–2569. Ding, Y., Ai, H.-w., Hoi, H., and Campbell, R.E. (2011) Forster resonance energy transfer-based biosensors for multiparameter ratiometric imaging of Ca2þ dynamics and caspase-3 activity in single cells. Analytical Chemistry, 83 (24), 9687–9693. Schifferer, M. and Griesbeck, O. (2012) A dynamic FRET reporter of gene expression improved by functional screening. Journal of the American Chemical Society, 134 (37), 15185–15188. Aoki, K., Komatsu, N., Hirata, E., Kamioka, Y., and Matsuda, M. (2012) Stable expression of FRET biosensors: a new light in cancer research. Cancer Science, 103 (4), 614–619. Sato, M. (2006) Imaging molecular events in single living cells. Analytical and Bioanalytical Chemistry, 386 (3), 435–443. Yoshiki, S., Matsunaga-Udagawa, R., Aoki, K., Kamioka, Y., Kiyokawa, E., and Matsuda, M. (2010) Ras and calcium signaling pathways converge at Raf1 via the Shoc2 scaffold protein. Molecular Biology of the Cell, 21 (6), 1088–1096. Albertazzi, L., Arosio, D., Marchetti, L., Ricci, F., and Beltram, F. (2009) Quantitative FRET analysis with the E(0) GFP-mCherry fluorescent protein pair. Photochemistry and Photobiology, 85 (1), 287–297. Shcherbakova, D.M., Hink, M.A., Joosen, L., Gadella, T.W.J., and Verkhusha, V.V. (2012) An orange fluorescent protein with a large Stokes shift for single-excitation multicolor FCCS and FRET imaging. Journal of the American Chemical Society, 134 (18), 7913–7923. Piljic, A. and Schultz, C. (2008) Analysis of protein complex hierarchy in living cells. ACS Chemical Biology, 3 (12), 749–755. Carlson, H.J. and Campbell, R.E. (2009) Genetically encoded FRET-based j249 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 243 244 245 246 247 248 249 250 251 biosensors for multiparameter fluorescence imaging. Current Opinion in Biotechnology, 20 (1), 19–27. Pires-Alves, M., Ho, M., Aberle, K.K., Janda, K.D., and Wilson, B.A. (2009) Tandem fluorescent proteins as enhanced FRET-based substrates for botulinum neurotoxin activity. Toxicon, 53 (4), 392–399. Bourgeois, D. and Adam, V. (2012) Reversible photoswitching in fluorescent proteins: a mechanistic view. IUBMB Life, 64 (6), 482–491. Voliani, V., Bizzarri, R., Nifosi, R., Abbruzzetti, S., Grandi, E., Viappiani, C., and Beltram, F. (2008) Cis–trans photoisomerization of fluorescent-protein chromophores. The Journal of Physical Chemistry. B, 112 (34), 10714–10722. Abbandonato, G., Signore, G., Nifosi, R., Voliani, V., Bizzarri, R., and Beltram, F. (2011) Cis–trans photoisomerization properties of GFP chromophore analogs. European Biophysics Journal with Biophysics Letters, 40 (11), 1205–1214. Grotjohann, T., Testa, I., Leutenegger, M., Bock, H., Urban, N.T., Lavoie-Cardinal, F., Willig, K.I., Eggeling, C., Jakobs, S., and Hell, S.W. (2011) Diffraction-unlimited all-optical imaging and writing with a photochromic GFP. Nature, 478 (7368), 204–208. Subach, F.V., Zhang, L., Gadella, T.W.J., Gurskaya, N.G., Lukyanov, K.A., and Verkhusha, V.V. (2010) Red fluorescent protein with reversibly photoswitchable absorbance for photochromic FRET. Chemistry & Biology, 17 (7), 745–755. Ando, R., Mizuno, H., and Miyawaki, A. (2004) Regulated fast nucleocytoplasmic shuttling observed by reversible protein highlighting. Science, 306 (5700), 1370–1373. Stiel, A.C., Andresen, M., Bock, H., Hilbert, M., Schilde, J., Schoenle, A., Eggeling, C., Egner, A., Hell, S.W., and Jakobs, S. (2008) Generation of monomeric reversibly switchable red fluorescent proteins for far-field fluorescence nanoscopy. Biophysical Journal, 95 (6), 2989–2997. Rice, K.G. (2001) Application of fluorescence resonance energy transfer to 252 253 254 255 256 257 258 259 260 analyze carbohydrates. Analytical Biochemistry, 297 (2), 117–122. Hoffmann, C., Gaietta, G., Bunemann, M., Adams, S.R., Oberdorff-Maass, S., Behr, B., Vilardaga, J.P., Tsien, R.Y., Eisman, M.H., and Lohse, M.J. (2005) A FlAsH-based FRET approach to determine G protein-coupled receptor activation in living cells. Nature Methods, 2 (3), 171–176. Neilson, J.A.D. and Durnford, D.G. (2010) Structural and functional diversification of the light-harvesting complexes in photosynthetic eukaryotes. Photosynthesis Research, 106 (1–2), 57–71. Glazer, A.N. and Stryer, L. (1984) Phycofluor probes. Trends in Biochemical Sciences, 9 (10), 423–427. Colyer, C., Kinkade, C., Viskari, P., and Landers, J. (2005) Analysis of cyanobacterial pigments and proteins by electrophoretic and chromatographic methods. Analytical and Bioanalytical Chemistry, 382 (3), 559–569. Sekar, S. and Chandramohan, M. (2008) Phycobiliproteins as a commodity: trends in applied research, patents and commercialization. Journal of Applied Phycology, 20 (2), 113–136. Schmidt, M.K. and Mackowski, S. (2011) Theoretical studies of excitation dynamics in a peridinin–chlorophyll– protein coupled to a metallic nanoparticle. Central European Journal of Physics, 9 (2), 562–569. Morseman, J.P., Moss, M.W., Zoha, S.J., and Allnutt, F.C.T. (1999) PBXL-1: a new fluorochrome applied to detection of proteins on membranes. Biotechniques, 26 (3), 559–563. Zoha, S.J., Ramnarain, S., Morseman, J. P., Moss, M.W., Allnutt, F.C.T., Rogers, Y.H., and Harvey, B. (1999) PBXL fluorescent dyes for ultrasensitive direct detection. Journal of Fluorescence, 9 (3), 197–208. Saraswat, S., Desireddy, A., Zheng, D., Guo, L., Lu, H.P., Bigioni, T.P., and Isailovic, D. (2011) Energy transfer from fluorescent proteins to metal nanoparticles. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 115 (35), 17587–17593. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 250 261 Schmitt, F.J., Maksimov, E.G., Haetti, P., 262 263 264 265 266 267 268 269 Weissenborn, J., Jeyasangar, V., Razjivin, A.P., Paschenko, V.Z., Friedrich, T., and Renger, G. (2012) Coupling of different isolated photosynthetic light harvesting complexes and CdSe/ZnS nanocrystals via Forster resonance energy transfer. Biochimica et Biophysica Acta, 1817 (8), 1461–1470. Gabourdes, M., Bourgine, V., Mathis, G., Bazin, H., and Alpha-Bazin, W. (2004) A homogeneous time-resolved fluorescence detection of telomerase activity. Analytical Biochemistry, 333 (1), 105–113. Dams, G., Van Acker, K., Gustin, E., Vereycken, I., Bunkens, L., Holemans, P., Smeulders, L., Clayton, R., Ohagen, A., and Hertogs, K. (2007) A time-resolved fluorescence assay to identify smallmolecule inhibitors of HIV-1 fusion. Journal of Biomolecular Screening, 12 (6), 865–874. Kota, S., Scampavia, L., Spicer, T., Beeler, A.B., Takahashi, V., Snyder, J.K., Porco, Jr., J.A., Hodder, P., and Strosberg, A.D. (2010) A time-resolved fluorescence-resonance energy transfer assay for identifying inhibitors of hepatitis C virus core dimerization. Assay and Drug Development Technologies, 8 (1), 96–105. Roh, Y.H., Ruiz, R.C.H., Peng, S., Lee, J.B., and Luo, D. (2011) Engineering DNA-based functional materials. Chemical Society Reviews, 40 (12), 5730–5744. Torring, T., Voigt, N.V., Nangreave, J., Yan, H., and Gothelf, K.V. (2011) DNA origami: a quantum leap for self-assembly of complex structures. Chemical Society Reviews, 40 (12), 5636–5646. Ke, Y., Ong, L.L., Shih, W.M., and Yin, P. (2012) Three-dimensional structures selfassembled from DNA bricks. Science, 338 (6111), 1177–1183. Li, Y.G., Tseng, Y.D., Kwon, S.Y., D’Espaux, L., Bunch, J.S., McEuen, P.L., and Luo, D. (2004) Controlled assembly of dendrimer-like DNA. Nature Materials, 3 (1), 38–42. Albinsson, B., Hannestad, J.K., and Boerjesson, K. (2012) Functionalized DNA nanostructures for light harvesting and charge separation. 270 271 272 273 274 275 276 277 Coordination Chemistry Reviews, 256 (21–22), 2399–2413. Teo, Y.N. and Kool, E.T. (2009) Polyfluorophore excimers and exciplexes as FRET donors in DNA. Bioconjugate Chemistry, 20 (12), 2371–2380. Kashida, H., Sekiguchi, K., Liang, X., and Asanuma, H. (2010) Accumulation of fluorophores into DNA duplexes to mimic the properties of quantum dots. Journal of the American Chemical Society, 132 (17), 6223–6230. Kumar, C.V. and Duff, Jr., M.R. (2009) DNA-based supramolecular artificial light harvesting complexes. Journal of the American Chemical Society, 131 (44), 16024–16026. Benvin, A.L., Creeger, Y., Fisher, G.W., Ballou, B., Waggoner, A.S., and Armitage, B.A. (2007) Fluorescent DNA nanotags: supramolecular fluorescent labels based on intercalating dye arrays assembled on nanostructured DNA templates. Journal of the American Chemical Society, 129 (7), 2025–2034. Oezhalici-Uenal, H. and Armitage, B.A. (2009) Fluorescent DNA nanotags based on a self-assembled DNA tetrahedron. ACS Nano, 3 (2), 425–433. Stadler, A.L., Delos Santos, J.O., Stensrud, E.S., Dembska, A., Silva, G.L., Liu, S., Shank, N.I., Kunttas-Tatli, E., Sobers, C.J., Gramlich, P.M.E., Care, T., Peteanu, L.A., McCartney, B.M., and Armitage, B.A. (2011) Fluorescent DNA nanotags featuring covalently attached intercalating dyes: synthesis, antibody conjugation, and intracellular imaging. Bioconjugate Chemistry, 22 (8), 1491–1502. Hall, M.P., Unch, J., Binkowski, B.F., Valley, M.P., Butler, B.L., Wood, M.G., Otto, P., Zimmerman, K., Vidugiris, G., Machleidt, T., Robers, M.B., Benink, H.A., Eggers, C.T., Slater, M.R., Meisenheimer, P.L., Klaubert, D.H., Fan, F., Encell, L.P., and Wood, K.V. (2012) Engineered luciferase reporter from a deep sea shrimp utilizing a novel imidazopyrazinone substrate. ACS Chemical Biology, 7 (11), 1848–1857. Wilson, T. and Hastings, J.W. (1998) Bioluminescence. Annual Review of Cell and Developmental Biology, 14, 197–230. j251 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 278 Roda, A., Guardigli, M., Michelini, E., and 279 280 281 282 283 284 285 286 287 Mirasoli, M. (2009) Bioluminescence in analytical chemistry and in vivo imaging. Trends in Analytical Chemistry, 28 (3), 307–322. Nakajima, Y. and Ohmiya, Y. (2010) Bioluminescence assays: multicolor luciferase assay, secreted luciferase assay and imaging luciferase assay. Expert Opinion on Drug Discovery, 5 (9), 835–849. Scott, D., Dikici, E., Ensor, M., and Daunert, S. (2011) Bioluminescence and its impact on bioanalysis. Annual Review of Analytical Chemistry, 4, 297–319. Kelkar, M. and De, A. (2012) Bioluminescence based in vivo screening technologies. Current Opinion in Pharmacology, 12 (5), 592–600. Roda, A. and Guardigli, M. (2012) Analytical chemiluminescence and bioluminescence: latest achievements and new horizons. Analytical and Bioanalytical Chemistry, 402 (1), 69–76. Algar, W.R., Tavares, A.J., and Krull, U.J. (2010) Beyond labels: a review of the application of quantum dots as integrated components of assays, bioprobes, and biosensors utilizing optical transduction. Analytica Chimica Acta, 673 (1), 1–25. Couturier, C. and Deprez, B. (2012) Setting up a bioluminescence resonance energy transfer high throughput screening assay to search for protein/ protein interaction inhibitors in mammalian cells. Frontiers in Endocrinology, 3, 100. Pfleger, K.D.G. and Eidne, K.A. (2006) Illuminating insights into protein–protein interactions using bioluminescence resonance energy transfer (BRET). Nature Methods, 3 (3), 165–174. Xia, Z. and Rao, J. (2009) Biosensing and imaging based on bioluminescence resonance energy transfer. Current Opinion in Biotechnology, 20 (1), 37–44. Alvarez-Curto, E., Pediani, J.D., and Milligan, G. (2010) Applications of fluorescence and bioluminescence resonance energy transfer to drug discovery at G protein coupled receptors. Analytical and Bioanalytical Chemistry, 398 (1), 167–180. 288 Lohse, M.J., Nuber, S., and Hoffmann, C. 289 290 291 292 293 294 295 296 (2012) Fluorescence/bioluminescence resonance energy transfer techniques to study G-protein-coupled receptor activation and signaling. Pharmacological Reviews, 64 (2), 299–336. Boute, N., Jockers, R., and Issad, T. (2002) The use of resonance energy transfer in high-throughput screening: BRET versus FRET. Trends in Pharmacological Sciences, 23 (8), 351–354. Perroy, J., Pontier, S., Charest, P.G., Aubry, M., and Bouvier, M. (2004) Realtime monitoring of ubiquitination in living cells by BRET. Nature Methods, 1 (3), 203–208. Sun, Y.-Q., Liu, J., Wang, P., Zhang, J., and Guo, W. (2012) D-luciferin analogues: a multicolor toolbox for bioluminescence imaging. Angewandte Chemie, International Edition, 51 (34), 8428–8430. Meroni, G., Rajabi, M., and Santaniello, E. (2009) D-luciferin, derivatives and analogues: synthesis and in vitro/in vivo luciferase-catalyzed bioluminescent activity. ARKIVOC, 2009, 265–288. Yang, J.W. and Thomason, D.B. (1993) An easily synthesized, photolyzable luciferase substrate for in vivo luciferase activity measurement. Biotechniques, 15 (5), 848–850. Van de Bittner, G.C., Dubikovskaya, E.A., Bertozzi, C.R., and Chang, C.J. (2010) In vivo imaging of hydrogen peroxide production in a murine tumor model with a chemoselective bioluminescent reporter. Proceedings of the National Academy of Sciences of the United States of America, 107 (50), 21316–21321. Inouye, S. and Shimomura, O. (1997) The use of Renilla luciferase, Oplophorus luciferase, and apoaequorin as bioluminescent reporter protein in the presence of coelenterazine analogues as substrate. Biochemical and Biophysical Research Communications, 233 (2), 349–353. Wu, C., Nakamura, H., Murai, A., and Shimomura, O. (2001) Chemi- and bioluminescence of coelenterazine analogues with a conjugated group at the C-8 position. Tetrahedron Letters, 42 (16), 2997–3000. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 252 297 Li, F., Yu, J., Zhang, Z., Cui, Z., Wang, D., 298 299 300 301 302 303 304 305 306 Wei, H., and Zhang, X.-E. (2012) Buffer enhanced bioluminescence resonance energy transfer sensor based on Gaussia luciferase for in vitro detection of protease. Analytica Chimica Acta, 724, 104–110. Francisco, W.A., Abusoud, H.M., Baldwin, T.O., and Raushel, F.M. (1993) Interaction of bacterial luciferase with aldehyde substrates and inhibitors. The Journal of Biological Chemistry, 268 (33), 24734–24741. Baldwin, T.O., Christopher, J.A., Raushel, F.M., Sinclair, J.F., Ziegler, M.M., Fisher, A.J., and Rayment, I. (1995) Structure of bacterial luciferase. Current Opinion in Structural Biology, 5 (6), 798–809. Wood, K.V. (1998) The chemistry of bioluminescent reporter assays. Promega Notes, 65, 7. Airth, R.L., Rhodes, W.C., and McElroy, W.D. (1958) The function of coenzyme A in luminescence. Biochimica et Biophysica Acta, 27 (3), 519–532. Dragulescu-Andrasi, A., Chan, C.T., De, A., Massoud, T.F., and Gambhir, S.S. (2011) Bioluminescence resonance energy transfer (BRET) imaging of protein–protein interactions within deep tissues of living subjects. Proceedings of the National Academy of Sciences of the United States of America, 108 (29), 12060–12065. Dacres, H., Michie, M., and Trowell, S.C. (2012) Comparison of enhanced bioluminescence energy transfer donors for protease biosensors. Analytical Biochemistry, 424 (2), 206–210. Bryan, B.J. and Szent-Gyorgyi, C. (2001) Luciferases, fluorescent proteins, nucleic acids encoding the luciferases and fluorescent proteins and the use thereof in diagnostics, high throughput screening and novelty items. U.S. Patent No. 6,436,682. Verhaegen, M. and Christopoulos, T.K. (2002) Recombinant Gaussia luciferase. Overexpression, purification, and analytical application of a bioluminescent reporter for DNA hybridization. Analytical Chemistry, 74 (17), 4378–4385. Kim, S.B., Suzuki, H., Sato, M., and Tao, H. (2011) Superluminescent variants of 307 308 309 310 311 312 313 314 315 marine luciferases for bioassays. Analytical Chemistry, 83 (22), 8732–8740. Tannous, B.A., Kim, D.E., Fernandez, J.L., Weissleder, R., and Breakefield, X.O. (2005) Codon-optimized Gaussia luciferase cDNA for mammalian gene expression in culture and in vivo. Molecular Therapy, 11 (3), 435–443. Chiesa, A., Rapizzi, E., Tosello, V., Pinton, P., de Virgilio, M., Fogarty, K.E., and Rizzuto, R. (2001) Recombinant aequorin and green fluorescent protein as valuable tools in the study of cell signalling. Biochemical Journal, 355, 1–12. Kendall, J.M. and Badminton, M.N. (1998) Aequorea victoria bioluminescence moves into an exciting new era. Trends in Biotechnology, 16 (5), 216–224. Alvarez, J. and Montero, M. (2002) Measuring Ca2þ in the endoplasmic reticulum with aequorin. Cell Calcium, 32 (5–6), 251–260. Gorokhovatsky, A.Y., Rudenko, N.V., Marchenkov, V.V., Skosyrev, V.S., Arzhanov, M.A., Burkhardt, N., Zakharov, M.V., Semisotnov, G.V., Vinokurov, L.M., and Alakhov, Y.B. (2003) Homogeneous assay for biotin based on Aequorea victoria bioluminescence resonance energy transfer system. Analytical Biochemistry, 313 (1), 68–75. De, A. and Gambhir, S.S. (2005) Noninvasive imaging of protein–protein interactions from live cells and living subjects using bioluminescence resonance energy transfer. FASEB Journal, 19 (12), 2017–2019. De, A., Ray, P., Loening, A.M., and Gambhir, S.S. (2009) BRET3: a red-shifted bioluminescence resonance energy transfer (BRET)-based integrated platform for imaging protein–protein interactions from single live cells and living animals. FASEB Journal, 23 (8), 2702–2709. Ayoub, M.A. and Pfleger, K.D.G. (2010) Recent advances in bioluminescence resonance energy transfer technologies to study GPCR heteromerization. Current Opinion in Pharmacology, 10 (1), 44–52. Walls, Z.F. and Gambhir, S.S. (2008) BRET-based method for detection of specific RNA species. Bioconjugate Chemistry, 19 (1), 178–184. j253 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 316 Ke, D. and Tu, S.-C. (2011) Activities, 317 318 319 320 321 322 323 324 325 kinetics and emission spectra of bacterial luciferase-fluorescent protein fusion enzymes. Photochemistry and Photobiology, 87 (6), 1346–1353. Carriba, P., Navarro, G., Ciruela, F., Ferre, S., Casado, V., Agnati, L., Cortes, A., Mallol, J., Fuxe, K., Canela, E.I., Lluis, C., and Franco, R. (2008) Detection of heteromerization of more than two proteins by sequential BRET–FRET. Nature Methods, 5 (8), 727–733. Branchini, B.R., Rosenberg, J.C., Ablamsky, D.M., Taylor, K.P., Southworth, T.L., and Linder, S.J. (2011) Sequential bioluminescence resonance energy transfer-fluorescence resonance energy transfer-based ratiometric protease assays with fusion proteins of firefly luciferase and red fluorescent protein. Analytical Biochemistry, 414 (2), 239–245. So, M.K., Xu, C.J., Loening, A.M., Gambhir, S.S., and Rao, J.H. (2006) Selfilluminating quantum dot conjugates for in vivo imaging. Nature Biotechnology, 24 (3), 339–343. Alam, R., Fontaine, D.M., Branchini, B.R., and Maye, M.M. (2012) Designing quantum rods for optimized energy transfer with firefly luciferase enzymes. Nano Letters, 12 (6), 3251–3256. Kim, G.B. and Kim, Y.-P. (2012) Analysis of protease activity using quantum dots and resonance energy transfer. Theranostics, 2 (2), 127–138. Kumar, M., Zhang, D., Broyles, D., and Deo, S.K. (2011) A rapid, sensitive, and selective bioluminescence resonance energy transfer (BRET)-based nucleic acid sensing system. Biosensors & Bioelectronics, 30 (1), 133–139. Li, Q., Zhang, L., Li, J., and Lu, C. (2011) Nanomaterial-amplified chemiluminescence systems and their applications in bioassays. Trends in Analytical Chemistry, 30 (2), 401–413. Rodriguez-Orozco, A.R., Ruiz-Reyes, H., and Medina-Serriteno, N. (2010) Recent applications of chemiluminescence assays in clinical immunology. Mini-Reviews in Medicinal Chemistry, 10 (14), 1393–1400. Lara, F.J., Garcia-Campana, A.M., and Aaron, J.-J. (2010) Analytical applications 326 327 328 329 330 331 332 333 of photoinduced chemiluminescence in flow systems: a review. Analytica Chimica Acta, 679 (1–2), 17–30. Wang, H.-Q., Li, Y.-Q., Wang, J.-H., Xu, Q., Li, X.-Q., and Zhao, Y.-D. (2008) Influence of quantum dot’s quantum yield to chemiluminescent resonance energy transfer. Analytica Chimica Acta, 610 (1), 68–73. Huang, X. and Ren, J. (2012) Nanomaterial-based chemiluminescence resonance energy transfer: a strategy to develop new analytical methods. Trends in Analytical Chemistry, 40, 77–89. Freeman, R., Liu, X., and Willner, I. (2011) Chemiluminescent and chemiluminescence resonance energy transfer (CRET) detection of DNA, metal ions, and aptamer-substrate complexes using hemin/G-quadruplexes and CdSe/ ZnS quantum dots. Journal of the American Chemical Society, 133 (30), 11597–11604. Liu, X., Freeman, R., Golub, E., and Willner, I. (2011) Chemiluminescence and chemiluminescence resonance energy transfer (CRET) aptamer sensors using catalytic hemin/G-quadruplexes. ACS Nano, 5 (9), 7648–7655. Messeri, G., Orlandini, A., and Pazzagli, M. (1989) Luminescent immunoassay using isoluminol derivatives. Journal of Bioluminescence and Chemiluminescence, 4 (1), 154–158. Richards, S.J. and Wusteman, F.S. (1989) Alkali-liable luminol derivatives as labels for quantitative binding-studies with polyionic macromolecules. Journal of Bioluminescence and Chemiluminescence, 3 (4), 175–179. Diaz, A.N., Sanchez, F.G., and Garcia, J.A. G. (1996) Hydrogen peroxide assay by using enhanced chemiluminescence of the luminol-H2O2-horseradish peroxidase system: comparative studies. Analytica Chimica Acta, 327 (2), 161–165. Easton, P.M., Simmonds, A.C., Rakishev, A., Egorov, A.M., and Candeias, L.P. (1996) Quantitative model of the enhancement of peroxidase-induced luminol luminescence. Journal of the American Chemical Society, 118 (28), 6619–6624. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 254 334 Wang, J.N. and Ren, J.C. (2005) A 335 336 337 338 339 340 341 342 sensitive and rapid immunoassay for quantification of CA125 in human sera by capillary electrophoresis with enhanced chemiluminescence detection. Electrophoresis, 26 (12), 2402–2408. Akhavan-Tafti, H., DeSilva, R., Arghavani, Z., Eickholt, R.A., Handley, R.S., Schoenfelner, B.A., Sugioka, K., Sugioka, Y., and Schaap, A.P. (1998) Characterization of acridancarboxylic acid derivatives as chemiluminescent peroxidase substrates. Journal of Organic Chemistry, 63 (4), 930–937. Akhavan-Tafti, H., deSilva, R., Eickholt, R., Handley, R., Mazelis, M., and Sandison, M. (2003) Characterization of new fluorescent peroxidase substrates. Talanta, 60 (2–3), 345–354. Osman, A.M., Zomer, G., Laane, C., and Hilhorst, R. (2000) Comparative studies of the chemiluminescent horseradish peroxidase-catalysed peroxidation of acridan (GZ-11) and luminol reactions: effect of pH and scavengers of reactive oxygen species on the light intensity of these systems. Luminescence, 15 (3), 189–197. Schaap, A.P., Akhavan, H., and Romano, L.J. (1989) Chemiluminescent substrates for alkaline phosphatase: application to ultrasensitive enzyme-linked immunoassays and DNA probes. Clinical Chemistry, 35 (9), 1863–1864. Beck, S. and Koster, H. (1990) Application of dioxetane chemiluminescent probes to molecular biology. Analytical Chemistry, 62 (21), 2258–2270. Zhao, S., Huang, Y., Shi, M., Liu, R., and Liu, Y.-M. (2010) Chemiluminescence resonance energy transfer-based detection for microchip electrophoresis. Analytical Chemistry, 82 (5), 2036–2041. Zhang, S., Yan, Y., and Bi, S. (2009) Design of molecular beacons as signaling probes for adenosine triphosphate detection in cancer cells based on chemiluminescence resonance energy transfer. Analytical Chemistry, 81 (21), 8695–8701. Qin, G., Zhao, S., Huang, Y., Jiang, J., and Ye, F. (2012) Magnetic bead-sensingplatform-based chemiluminescence 343 344 345 346 347 348 349 350 351 resonance energy transfer and its immunoassay application. Analytical Chemistry, 84 (6), 2708–2712. Huang, X. and Ren, J. (2011) Gold nanoparticles based chemiluminescent resonance energy transfer for immunoassay of alpha fetoprotein cancer marker. Analytica Chimica Acta, 686 (1–2), 115–120. Bi, S., Zhang, J., Hao, S., Ding, C., and Zhang, S. (2011) Exponential amplification for chemiluminescence resonance energy transfer detection of MicroRNA in real samples based on a cross-catalyst strand-displacement network. Analytical Chemistry, 83 (10), 3696–3702. Golub, E., Niazov, A., Freeman, R., and Zatsepin, M., and Willner, I. (2012) Photoelectrochemical biosensors without external irradiation: probing enzyme activities and DNA sensing using hemin/ G-quadruplex-stimulated chemiluminescence resonance energy transfer (CRET) generation of photocurrents. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 116 (25), 13827–13834. Selvin, P.R. (2002) Principles and biophysical applications of lanthanidebased probes. Annual Review of Biophysics and Biomolecular Structure, 31, 275–302. Buenzli, J.-C.G. (2010) Lanthanide luminescence for biomedical analyses and imaging. Chemical Reviews, 110 (5), 2729–2755. Eliseeva, S.V. and Buenzli, J.-C.G. (2010) Lanthanide luminescence for functional materials and bio-sciences. Chemical Society Reviews, 39 (1), 189–227. Hemmila, I. and Laitala, V. (2005) Progress in lanthanides as luminescent probes. Journal of Fluorescence, 15 (4), 529–542. Kadjane, P., Starck, M., Camerel, F., Hill, D., Hildebrandt, N., Ziessel, R., and Charbonniere, L.J. (2009) Divergent approach to a large variety of versatile luminescent lanthanide complexes. Inorganic Chemistry, 48 (11), 4601–4603. Bourdolle, A., Allali, M., Mulatier, J.-C., Le Guennic, B., Zwier, J.M., Baldeck, P.L., Buenzli, J.-C.G., Andraud, C., Lamarque, j255 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 352 353 354 355 356 357 358 359 L., and Maury, O. (2011) Modulating the photophysical properties of azamacrocyclic europium complexes with charge-transfer antenna chromophores. Inorganic Chemistry, 50 (11), 4987–4999. Kupcho, K.R., Stafslien, D.K., DeRosier, T., Hallis, T.M., Ozers, M.S., and Vogel, K. W. (2007) Simultaneous monitoring of discrete binding events using dualacceptor terbium-based LRET. Journal of the American Chemical Society, 129 (44), 13372–13373. Kokko, T., Kokko, L., and Soukka, T. (2009) Terbium(III) chelate as an efficient donor for multiple-wavelength fluorescent acceptors. Journal of Fluorescence, 19 (1), 159–164. Vuojola, J., Hyppanen, I., Nummela, M., Kankare, J., and Soukka, T. (2011) Distance and temperature dependency in nonoverlapping and conventional Forster resonance energy-transfer. The Journal of Physical Chemistry. B, 115 (46), 13685– 13694. Vuojola, J., Lamminmaki, U., and Soukka, T. (2009) Resonance energy transfer from lanthanide chelates to overlapping and nonoverlapping fluorescent protein acceptors. Analytical Chemistry, 81 (12), 5033–5038. Rajapakse, H.E., Gahlaut, N., Mohandessi, S., Yu, D., Turner, J.R., and Miller, L.W. (2010) Time-resolved luminescence resonance energy transfer imaging of protein–protein interactions in living cells. Proceedings of the National Academy of Sciences of the United States of America, 107 (31), 13582–13587. Kim, B., Tarchevskaya, S.S., Eggel, A., Vogel, M., and Jardetzky, T.S. (2012) A time-resolved fluorescence resonance energy transfer assay suitable for highthroughput screening for inhibitors of immunoglobulin E-receptor interactions. Analytical Biochemistry, 431 (2), 84–89. Kokko, T., Liljenback, T., Peltola, M.T., Kokko, L., and Soukka, T. (2008) Homogeneous dual-parameter assay for prostate-specific antigen based on fluorescence resonance energy transfer. Analytical chemistry, 80 (24), 9763–9768. Martikkala, E., Veltel, S., Kirjavainen, J., Rozwandowicz-Jansen, A., Lamminmaki, 360 361 362 363 364 365 366 U., Hanninen, P., and Harma, H. (2011) Homogeneous single-label biochemical Ras activation assay using time-resolved luminescence. Analytical Chemistry, 83 (24), 9230–9233. Kupstat, A., Kumke, M.U., and Hildebrandt, N. (2011) Toward sensitive, quantitative point-of-care testing (POCT) of protein markers: miniaturization of a homogeneous time-resolved fluoroimmunoassay for prostate-specific antigen detection. Analyst, 136 (5), 1029–1035. Li, L.-L., Ge, P., Selvin, P.R., and Lu, Y. (2012) Direct detection of adenosine in undiluted serum using a luminescent aptamer sensor attached to a terbium complex. Analytical Chemistry, 84 (18), 7852–7856. Geißsler, D., Stufler, S., Lohmannsroben, H.-G., and Hildebrandt, N. (2013) Sixcolor time-resolved Forster resonance energy transfer for ultrasensitive multiplexed biosensing. Journal of the American Chemical Society, 135 (3), 1102–1109. Charbonniere, L.J. and Hildebrandt, N. (2008) Lanthanide complexes and quantum dots: a bright wedding for resonance energy transfer. European Journal of Inorganic Chemistry, 2008 (21), 3241–3251. Charbonniere, L.J., Hildebrandt, N., Ziessel, R.F., and Loehmannsroeben, H.G. (2006) Lanthanides to quantum dots resonance energy transfer in timeresolved fluoro-immunoassays and luminescence microscopy. Journal of the American Chemical Society, 128 (39), 12800–12809. Algar, W.R., Malanoski, A.P., Susumu, K., Stewart, M.H., Hildebrandt, N., and Medintz, I.L. (2012) Multiplexed tracking of protease activity using a single color of quantum dot vector and a time-gated Forster resonance energy transfer relay. Analytical Chemistry, 84 (22), 10136–10146. Algar, W.R., Wegner, D., Huston, A.L., Blanco-Canosa, J.B., Stewart, M.H., Armstrong, A., Dawson, P.E., Hildebrandt, N., and Medintz, I.L. (2012) Quantum dots as simultaneous acceptors Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 256 367 368 369 370 371 372 373 374 and donors in time-gated Forster resonance energy transfer relays: characterization and biosensing. Journal of the American Chemical Society, 134 (3), 1876–1891. Chen, M.-J., Wu, Y.-S., Lin, G.-F., Hou, J.Y., Li, M., and Liu, T.-C. (2012) Quantumdot-based homogeneous time-resolved fluoroimmunoassay of alpha-fetoprotein. Analytica Chimica Acta, 741, 100–105. Algar, W.R., Malonoski, A., Deschamps, J. R., Banco-Canosa, J.B., Susumu, K., Stewart, M.H., Johnson, B.J., Dawson, P. E., and Medintz, I.L. (2012) Proteolytic activity at quantum dot-conjugates: kinetic analysis reveals enhanced enzyme activity and localized interfacial “hopping”. Nano Letters, 12 (7), 3793–3802. Harma, H., Suhonen, R., Kololuoma, T., Karkkainen, A., Hara, M., and Hanninen, P. (2009) Thin solid europium(III) dye layers as donors in time-resolved fluorescence resonance energy transfer assays. Applied Surface Science, 255 (13–14), 6529–6534. Dumke, J.C., El-Zahab, B., Challa, S., Das, S., Chandler, L., Tolocka, M., Hayes, D.J., and Warner, I.M. (2010) Lanthanide-based luminescent NanoGUMBOS. Langmuir, 26 (19), 15599–15603. Zhang, D., Wang, X., Qiao, Z.-a., Tang, D., Liu, Y., and Huo, Q. (2010) Synthesis and characterization of novel lanthanide(III) complexes-functionalized mesoporous silica nanoparticles as fluorescent nanomaterials. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 114 (29), 12505–12510. Song, C., Ye, Z., Wang, G., Yuan, J., and Guan, Y. (2010) Core–shell nanoarchitectures: a strategy to improve the efficiency of luminescence resonance energy transfer. ACS Nano, 4 (9), 5389–5397. Pihlasalo, S., Pellonpera, L., Martikkala, E., Hanninen, P., and Harma, H. (2010) Sensitive fluorometric nanoparticle assays for cell counting and viability. Analytical Chemistry, 82 (22), 9282–9288. Soukka, T., Rantanen, T., and Kuningas, K. (2008) Photon upconversion in homogeneous fluorescence-based bioanalytical assays, in Fluorescence 375 376 377 378 379 380 381 382 Methods and Applications: Spectroscopy, Imaging, and Probes (ed. O.S. Wolfbeis), Wiley-Blackwell, pp. 188–200. Smith, A.M., Mancini, M.C., and Nie, S. (2009) Bioimaging: second window for in vivo imaging. Nature Nanotechnology, 4 (11), 710–711. Li, L.-L., Zhang, R., Yin, L., Zheng, K., Qin, W., Selvin, P.R., and Lu, Y. (2012) Biomimetic surface engineering of lanthanide-doped upconversion nanoparticles as versatile bioprobes. Angewandte Chemie, International Edition, 51 (25), 6121–6125. Wang, F., Banerjee, D., Liu, Y., Chen, X., and Liu, X. (2010) Upconversion nanoparticles in biological labeling, imaging, and therapy. Analyst, 135 (8), 1839–1854. Dong, N.-N., Pedroni, M., Piccinelli, F., Conti, G., Sbarbati, A., Enrique RamirezHernandez, J., Martinez Maestro, L., Carmen Iglesias-de la Cruz, M., SanzRodriguez, F., Juarranz, A., Chen, F., Vetrone, F., Capobianco, J.A., Garcia Sole, J., Bettinelli, M., Jaque, D., and Speghini, A. (2011) NIR-to-NIR two-photon excited CaF2: Tm3þ,Yb3þ nanoparticles: multifunctional nanoprobes for highly penetrating fluorescence bio-imaging. ACS Nano, 5 (11), 8665–8671. Bouzigues, C., Gacoin, T., and Alexandrou, A. (2011) Biological applications of rare-earth based nanoparticles. ACS Nano, 5 (11), 8488–8505. Zhang, C., Sun, L., Zhang, Y., and Yan, C. (2010) Rare earth upconversion nanophosphors: synthesis, functionalization and application as biolabels and energy transfer donors. Journal of Rare Earths, 28 (6), 807–819. Lin, M., Zhao, Y., Wang, S., Liu, M., Duan, Z., Chen, Y., Li, F., Xu, F., and Lu, T. (2012) Recent advances in synthesis and surface modification of lanthanidedoped upconversion nanoparticles for biomedical applications. Biotechnology Advances, 30 (6), 1551–1561. Huang, Q., Yu, J., Ma, E., and Lin, K. (2010) Synthesis and characterization of highly efficient near-infrared upconversion Sc3þ/Er3þ/Yb3þ tridoped j257 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 383 384 385 386 387 388 389 390 NaYF4. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 114 (10), 4719–4724. Yang, J., Shen, D., Li, X., Li, W., Fang, Y., Wei, Y., Yao, C., Tu, B., Zhang, F., and Zhao, D. (2012) One-step hydrothermal synthesis of carboxyl-functionalized upconversion phosphors for bioapplications. Chemistry: A European Journal, 18 (43), 13642–13650. Bednarkiewicz, A., Wawrzynczyk, D., Gagor, A., Kepinski, L., Kurnatowska, M., Krajczyk, L., Nyk, M., Samoc, M., and Strek, W. (2012) Giant enhancement of upconversion in ultra-small Er3þ/Yb3þ: NaYF4 nanoparticles via laser annealing. Nanotechnology, 23 (14), 145705. Corstjens, P.L.A.M., Li, S., Zuiderwijk, M., Kardos, K., Abrams, W.R., Niedbala, R.S., and Tanke, H.J. (2005) Infrared upconverting phosphors for bioassays. IEE Proceedings. Nanobiotechnology, 152 (2), 64–72. Ju, Q., Tu, D., Liu, Y., Zhu, H., and Chen, X. (2012) Lanthanide-doped inorganic nanocrystals as luminescent biolabels. Combinatorial Chemistry & High Throughput Screening, 15 (7), 580–594. Gu, J.-Q., Shen, J., Sun, L.-D., and Yan, C.H. (2008) Resonance energy transfer in steady-state and time-decay fluoroimmunoassays for lanthanide nanoparticles based on biotin and avidin affinity. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 112 (17), 6589–6593. Casanova, D., Giaume, D., Moreau, M., Martin, J.-L., Gacoin, T., Boilot, J.-P., and Alexandrou, A. (2007) Counting the number of proteins coupled to single nanoparticles. Journal of the American Chemical Society, 129 (42), 12592–12593. Di, W., Li, J., Shirahata, N., and Sakka, Y. (2010) An efficient and biocompatible fluorescence resonance energy transfer system based on lanthanide-doped nanoparticles. Nanotechnology, 21 (45), 455703. Liu, Y., Zhou, S., Tu, D., Chen, Z., Huang, M., Zhu, H., Ma, E., and Chen, X. (2012) Amine-functionalized lanthanide-doped zirconia nanoparticles: optical spectroscopy, time-resolved fluorescence 391 392 393 394 395 396 397 398 399 400 resonance energy transfer biodetection, and targeted imaging. Journal of the American Chemical Society, 134 (36), 15083–15090. Auzel, F. (2004) Upconversion and antiStokes processes with f and d ions in solids. Chemical Reviews, 104 (1), 139–173. Wang, G., Peng, Q., and Li, Y. (2011) Lanthanide-doped nanocrystals: synthesis, optical-magnetic properties, and applications. Accounts of Chemical Research, 44 (5), 322–332. Bednarkiewicz, A., Nyk, M., Samoc, M., and Strek, W. (2010) Up-conversion FRET from Er3þ/Yb3þ:NaYF4 nanophosphor to CdSe quantum dots. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 114 (41), 17535–17541. Peng, J., Wang, Y., Wang, J., Zhou, X., and Liu, Z. (2011) A new biosensor for glucose determination in serum based on upconverting fluorescence resonance energy transfer. Biosensors & Bioelectronics, 28 (1), 414–420. Mader, H.S. and Wolfbeis, O.S. (2010) Optical ammonia sensor based on upconverting luminescent nanoparticles. Analytical Chemistry, 82 (12), 5002–5004. Rantanen, T., Jarvenpaa, M.-L., Vuojola, J., Arppe, R., Kuningas, K., and Soukka, T. (2009) Upconverting phosphors in a dualparameter LRET-based hybridization assay. Analyst, 134 (8), 1713–1716. Vuojola, J., Riuttamaki, T., Kulta, E., Arppe, R., and Soukka, T. (2012) Fluorescence-quenching-based homogeneous caspase-3 activity assay using photon upconversion. Analytica Chimica Acta, 725, 67–73. Jiang, S. and Zhang, Y. (2010) Upconversion nanoparticle-based FRET system for study of siRNA in live cells. Langmuir, 26 (9), 6689–6694. Cheng, L., Yang, K., Shao, M., Lee, S.-T., and Liu, Z. (2011) Multicolor in vivo imaging of upconversion nanoparticles with emissions tuned by luminescence resonance energy transfer. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 115 (6), 2686–2692. Piszczek, G. (2006) Luminescent metal– ligand complexes as probes of macromolecular interactions and Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 258 401 402 403 404 405 406 407 408 409 biopolymer dynamics. Archives of Biochemistry and Biophysics, 453 (1), 54–62. Patra, M. and Gasser, G. (2012) Organometallic compounds: an opportunity for chemical biology? ChemBioChem, 13 (9), 1232–1252. Fernandez-Moreira, V., ThorpGreenwood, F.L., and Coogan, M.P. (2010) Application of d(6) transition metal complexes in fluorescence cell imaging. Chemical Communications, 46 (2), 186–202. Zhao, Q., Huang, C., and Li, F. (2011) Phosphorescent heavy-metal complexes for bioimaging. Chemical Society Reviews, 40 (5), 2508–2524. Thorp-Greenwood, F.L., Balasingham, R. G., and Coogan, M.P. (2012) Organometallic complexes of transition metals in luminescent cell imaging applications. Journal of Organometallic Chemistry, 714, 12–21. Balasingham, R.G., Thorp-Greenwood, F. L., Williams, C.F., Coogan, M.P., and Pope, S.J.A. (2012) Biologically compatible, phosphorescent dimetallic rhenium complexes linked through functionalized alkyl chains: syntheses, spectroscopic properties, and applications in imaging microscopy. Inorganic Chemistry, 51 (3), 1419–1426. Terpetschnig, E., Szmacinski, H., Malak, H., and Lakowicz, J.R. (1995) Metal– ligand complexes as a new class of longlived fluorophores for protein hydrodynamics. Biophysical Journal, 68 (1), 342–350. Youn, H.J., Terpetschnig, E., Szmacinski, H., and Lakowicz, J.R. (1995) Fluorescence energy transfer immunoassay based on a long-lifetime luminescent metal–ligand complex. Analytical Biochemistry, 232 (1), 24–30. Weh, J., Duerkop, A., and Wolfbeis, O.S. (2007) A resonance energy transfer immunoassay based on a thiol-reactive ruthenium donor dye and a longwaveemitting acceptor. ChemBioChem, 8 (1), 122–128. von Bultzingslowen, C., McEvoy, A.K., McDonagh, C., and MacCraith, B.D. (2003) Lifetime-based optical sensor for 410 411 412 413 414 415 416 417 high-level pCO(2) detection employing fluorescence resonance energy transfer. Analytica Chimica Acta, 480 (2), 275–283. Mareeswaran, P.M., Maheshwaran, D., Babu, E., and Rajagopal, S. (2012) Binding and fluorescence resonance energy transfer (FRET) of ruthenium(II)bipyridine-calixarene system with proteins-experimental and docking studies. Journal of Fluorescence, 22 (5), 1345–1356. Kainmueller, E.K. and Bannwarth, W. (2006) A new robust and highly sensitive FRET donor–acceptor pair: synthesis, characterization, and application in a thrombin assay. Helvetica Chimica Acta, 89 (12), 3056–3070. Gehne, S., Flehr, R., Kienzler, A.A.N., Berg, M., Bannwarth, W., and Kumke, M.U. (2012) Dye dynamics in threecolor FRET samples. The Journal of Physical Chemistry. B, 116 (35), 10798–10806. Kienzler, A., Flehr, R., Kramer, R.A., Gehne, S., Kumke, M.U., and Bannwarth, W. (2011) Novel three-color FRET tool box for advanced protein and DNA analysis. Bioconjugate Chemistry, 22 (9), 1852–1863. McLaurin, E.J., Greytak, A.B., Bawendi, M.G., and Nocera, D.G. (2009) Twophoton absorbing nanocrystal sensors for ratiometric detection of oxygen. Journal of the American Chemical Society, 131 (36), 12994–13001. Stewart, M.H., Huston, A.L., Scott, A.M., Efros, A.L., Melinger, J.S., Gemmill, K.B., Trammell, S.A., Blanco-Canosa, J.B., Dawson, P.E., and Medintz, I.L. (2012) Complex Forster energy transfer interactions between semiconductor quantum dots and a redox-active osmium assembly. ACS Nano, 6 (6), 5330–5347. Shiu, H.-Y., Wong, M.-K., and Che, C.-M. (2011) “Turn-on” FRET-based luminescent iridium(III) probes for the detection of cysteine and homocysteine. Chemical Communications, 47 (15), 4367–4369. Lee, A.J., Ensign, A.A., Krauss, T.D., and Bren, K.L. (2010) Zinc porphyrin as a donor for FRET in Zn(II)cytochrome c. Journal of the American Chemical Society, 132 (6), 1752–1753. j259 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 418 Chung, C.Y.-S. and Yam, V.W.-W. (2011) 419 420 421 422 423 424 425 426 Induced self-assembly and Forster resonance energy transfer studies of alkynylplatinum(II) terpyridine complex through interaction with water-soluble poly(phenylene ethynylene sulfonate) and the proof-of-principle demonstration of this two-component ensemble for selective label-free detection of human serum albumin. Journal of the American Chemical Society, 133 (46), 18775–18784. Oh, E., Delehanty, J.B., Sapsford, K.E., Susumu, K., Goswami, R., BlancoCanosa, J.B., Dawson, P.E., Granek, J., Shoff, M., Zhang, Q., Goering, P.L., Huston, A., and Medintz, I.L. (2011) Cellular uptake and fate of PEGylated gold nanoparticles is dependent on both cellpenetration peptides and particle size. ACS Nano, 5 (8), 6434–6448. Jain, P.K., Huang, X., El-Sayed, I.H., and El-Sayad, M.A. (2007) Review of some interesting surface plasmon resonanceenhanced properties of noble metal nanoparticles and their applications to biosystems. Plasmonics, 2 (3), 107–118. Deng, W. and Goldys, E.M. (2012) Plasmonic approach to enhanced fluorescence for applications in biotechnology and the life sciences. Langmuir, 28 (27), 10152–10163. Han, B. and Wang, E. (2012) DNAtemplated fluorescent silver nanoclusters. Analytical and Bioanalytical Chemistry, 402 (1), 129–138. Zheng, J., Nicovich, P.R., and Dickson, R. M. (2007) Highly fluorescent noble-metal quantum dots. Annual Review of Physical Chemistry, 58, 409–431. Zheng, J., Zhou, C., Yu, M., and Liu, J. (2012) Different sized luminescent gold nanoparticles. Nanoscale, 4 (14), 4073–4083. Daniel, M.C. and Astruc, D. (2004) Gold nanoparticles: assembly, supramolecular chemistry, quantum-size-related properties, and applications toward biology, catalysis, and nanotechnology. Chemical Reviews, 104 (1), 293–346. Radwan, S.H. and Azzazy, H.M.E. (2009) Gold nanoparticles for molecular diagnostics. Expert Review of Molecular Diagnostics, 9 (5), 511–524. 427 Diez, I. and Ras, R.H.A. (2011) 428 429 430 431 432 433 434 435 436 437 Fluorescent silver nanoclusters. Nanoscale, 3 (5), 1963–1970. Aslan, K., Gryczynski, I., Malicka, J., Matveeva, E., Lakowicz, J.R., and Geddes, C.D. (2005) Metal-enhanced fluorescence: an emerging tool in biotechnology. Current Opinion in Biotechnology, 16 (1), 55–62. Aslan, K., Lakowicz, J.R., and Geddes, C. D. (2005) Metal-enhanced fluorescence using anisotropic silver nanostructures: critical progress to date. Analytical and Bioanalytical Chemistry, 382 (4), 926–933. Polte, J., Ahner, T.T., Delissen, F., Sokolov, S., Emmerling, F., Thuenemann, A.F., and Kraehnert, R. (2010) Mechanism of gold nanoparticle formation in the classical citrate synthesis method derived from coupled in situ XANES and SAXS evaluation. Journal of the American Chemical Society, 132 (4), 1296–1301. Grzelczak, M., Perez-Juste, J., Mulvaney, P., and Liz-Marzan, L.M. (2008) Shape control in gold nanoparticle synthesis. Chemical Society Reviews, 37 (9), 1783–1791. Hu, M., Chen, J., Li, Z.-Y., Au, L., Hartland, G.V., Li, X., Marquez, M., and Xia, Y. (2006) Gold nanostructures: engineering their plasmonic properties for biomedical applications. Chemical Society Reviews, 35 (11), 1084–1094. Pastoriza-Santos, I., Alvarez-Puebla, R.A., and Liz-Marzan, L.M. (2010) Synthetic routes and plasmonic properties of noble metal nanoplates. European Journal of Inorganic Chemistry, 2010 (27), 4288–4297. Nair, L.S. and Laurencin, C.T. (2007) Silver nanoparticles: synthesis and therapeutic applications. Journal of Biomedical Nanotechnology, 3 (4), 301–316. Sintubin, L., Verstraete, W., and Boon, N. (2012) Biologically produced nanosilver: current state and future perspectives. Biotechnology and Bioengineering, 109 (10), 2422–2436. Hakkinen, H. (2012) The gold–sulfur interface at the nanoscale. Nature Chemistry, 4 (6), 443–455. Sperling, R.A. and Parak, W.J. (2010) Surface modification, functionalization and bioconjugation of colloidal inorganic Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 260 438 439 440 441 442 443 444 445 446 nanoparticles. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 368 (1915), 1333–1383. Maali, A., Cardinal, T., and TreguerDelapierre, M. (2003) Intrinsic fluorescence from individual silver nanoparticles. Physica E: Low-Dimensional Systems & Nanostructures, 17 (1–4), 559–560. Petty, J.T., Zheng, J., Hud, N.V., and Dickson, R.M. (2004) DNA-templated Ag nanocluster formation. Journal of the American Chemical Society, 126 (16), 5207–5212. Evanoff, D.D. and Chumanov, G. (2005) Synthesis and optical properties of silver nanoparticles and arrays. ChemPhysChem, 6 (7), 1221–1231. Felix, C., Sieber, C., Harbich, W., Buttet, J., Rabin, I., Schulze, W., and Ertl, G. (1999) Fluorescence and excitation spectra of Ag-4 in an argon matrix. Chemical Physics Letters, 313 (1–2), 105–109. Felix, C., Sieber, C., Harbich, W., Buttet, J., Rabin, I., Schulze, W., and Ertl, G. (2001) Ag-8 fluorescence in argon. Physical Review Letters, 86 (14), 2992–2995. Gao, J.P., Fu, J., Lin, C.K., Lin, J., Han, Y. C., Yu, X., and Pan, C.Y. (2004) Formation and photoluminescence of silver nanoparticles stabilized by a two-armed polymer with a crown ether core. Langmuir, 20 (22), 9775–9779. Dong, J., Zheng, H., Shao, H., Zhang, Z., Liu, G., and Qu, S.-X. (2011) Surface enhanced fluorescence of physically polished nanostructured metal surface: the effect of the native oxide layer. Journal of Nanoscience and Nanotechnology, 11 (11), 9626–9630. Chowdhury, S., Wu, Z., Jaquins-Gerstl, A., Liu, S., Dembska, A., Armitage, B.A., Jin, R., and Peteanu, L.A. (2011) Wavelength dependence of the fluorescence quenching efficiency of nearby dyes by gold nanoclusters and nanoparticles: the roles of spectral overlap and particle size. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 115 (41), 20105–20112. Pustovit, V.N. and Shahbazyan, T.V. (2012) Fluorescence quenching near small metal 447 448 449 450 451 452 453 454 455 nanoparticles. Journal of Chemical Physics, 136 (20), 204701. Craighead, H.G. and Glass, A.M. (1981) Optical-absorption of small metal particles with absorbed dye coats. Optics Letters, 6 (5), 248–250. Cohanoschi, I. and Hernandez, F.E. (2005) Surface plasmon enhancement of two- and three-photon absorption of Hoechst 33 258 dye in activated gold colloid solution. The Journal of Physical Chemistry. B, 109 (30), 14506–14512. Wang, D.S. and Kerker, M. (1982) Absorption and luminescence of dyecoated silver and gold particles. Physical Review. B, Condensed Matter and Materials Physics, 25 (4), 2433–2449. Zhang, J. and Lakowicz, J.R. (2005) Enhanced luminescence of phenylphenanthridine dye on aggregated small silver nanoparticles. The Journal of Physical Chemistry. B, 109 (18), 8701–8706. Zhang, J., Malicka, J., Gryczynski, I., and Lakowicz, J.R. (2005) Surfaceenhanced fluorescence of fluoresceinlabeled oligonucleotides capped on silver nanoparticles. The Journal of Physical Chemistry. B, 109 (16), 7643–7648. Lukomska, J., Malicka, J., Gryczynski, I., Leonenko, Z., and Lakowicz, J.R. (2005) Fluorescence enhancement of fluorophores tethered to different sized silver colloids deposited on glass substrate. Biopolymers, 77 (1), 31–37. Malicka, J., Gryczynski, I., Fang, J., Kusba, J., and Lakowicz, J.R. (2003) Increased resonance energy transfer between fluorophores bound to DNA in proximity to metallic silver particles. Analytical Biochemistry, 315 (2), 160–169. Malicka, J., Gryczynski, I., Fang, J.Y., and Lakowicz, J.R. (2003) Fluorescence spectral properties of cyanine dye-labeled DNA oligomers on surfaces coated with silver particles. Analytical Biochemistry, 317 (2), 136–146. Malicka, J., Gryczynski, I., Geddes, C.D., and Lakowicz, J.R. (2003) Metal-enhanced emission from indocyanine green: a new approach to in vivo imaging. Journal of Biomedical Optics, 8 (3), 472–478. j261 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 456 Malicka, J., Gryczynski, I., Gryczynski, Z., 457 458 459 460 461 462 463 and Lakowicz, J.R. (2003) Effects of fluorophore-to-silver distance on the emission of cyanine-dye-labeled oligonucleotides. Analytical Biochemistry, 315 (1), 57–66. Eustis, S. and El-Sayed, M.A. (2006) Why gold nanoparticles are more precious than pretty gold: noble metal surface plasmon resonance and its enhancement of the radiative and nonradiative properties of nanocrystals of different shapes. Chemical Society Reviews, 35 (3), 209–217. Swierczewska, M., Lee, S., and Chen, X. (2011) The design and application of fluorophore-gold nanoparticle activatable probes. Physical Chemistry Chemical Physics, 13 (21), 9929–9941. Kang, K.A., Wang, J., Jasinski, J.B., and Achilefu, S. (2011) Fluorescence manipulation by gold nanoparticles: from complete quenching to extensive enhancement. Journal of Nanobiotechnology, 9, 16. Dulkeith, E., Ringler, M., Klar, T.A., Feldmann, J., Javier, A.M., and Parak, W.J. (2005) Gold nanoparticles quench fluorescence by phase induced radiative rate suppression. Nano Letters, 5 (4), 585–589. Yun, C.S., Javier, A., Jennings, T., Fisher, M., Hira, S., Peterson, S., Hopkins, B., Reich, N.O., and Strouse, G.F. (2005) Nanometal surface energy transfer in optical rulers, breaking the FRET barrier. Journal of the American Chemical Society, 127 (9), 3115–3119. Pons, T., Medintz, I.L., Sapsford, K.E., Higashiya, S., Grimes, A.F., English, D.S., and Mattoussi, H. (2007) On the quenching of semiconductor quantum dot photoluminescence by proximal gold nanoparticles. Nano Letters, 7 (10), 3157–3164. Zhang, X., Marocico, C.A., Lunz, M., Gerard, V.A., Gun’ko, Y.K., Lesnyak, V., Gaponik, N., Susha, A.S., Rogach, A.L., and Bradley, A.L. (2012) Wavelength, concentration, and distance dependence of nonradiative energy transfer to a plane of gold nanoparticles. ACS Nano, 6 (10), 9283–9290. 464 Mandal, S., Ghatak, C., Rao, V.G., Ghosh, 465 466 467 468 469 470 471 472 S., and Sarkar, N. (2012) Pluronic micellar aggregates loaded with gold nanoparticles (Au NPs) and fluorescent dyes: a study of controlled nanometal surface energy transfer. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 116 (9), 5585–5597. Chatterjee, S., Lee, J.B., Valappil, N.V., Luo, D., and Menon, V.M. (2011) Investigating the distance limit of a metal nanoparticle based spectroscopic ruler. Biomedical Optics Express, 2 (6), 1727–1733. Dubertret, B., Calame, M., and Libchaber, A.J. (2001) Single-mismatch detection using gold-quenched fluorescent oligonucleotides. Nature Biotechnology, 19 (4), 365–370. Maxwell, D.J., Taylor, J.R., and Nie, S.M. (2002) Self-assembled nanoparticle probes for recognition and detection of biomolecules. Journal of the American Chemical Society, 124 (32), 9606–9612. Jennings, T.L., Schlatterer, J.C., Singh, M. P., Greenbaum, N.L., and Strouse, G.F. (2006) NSET molecular beacon analysis of hammerhead RNA substrate binding and catalysis. Nano Letters, 6 (7), 1318–1324. Huang, K. and Marti, A.A. (2012) Recent trends in molecular beacon design and applications. Analytical and Bioanalytical Chemistry, 402 (10), 3091–3102. Bunz, U.H.F. and Rotello, V.M. (2010) Gold nanoparticle-fluorophore complexes: sensitive and discerning “noses” for biosystems sensing. Angewandte Chemie, International Edition, 49 (19), 3268–3279. Guirgis, B.S.S., Sae Cunha, C., Gomes, I., Cavadas, M., Silva, I., Doria, G., Blatch, G. L., Baptista, P.V., Pereira, E., Azzazy, H. M.E., Mota, M.M., Prudencio, M., and Franco, R. (2012) Gold nanoparticle-based fluorescence immunoassay for malaria antigen detection. Analytical and Bioanalytical Chemistry, 402 (3), 1019–1027. Song, S., Liang, Z., Zhang, J., Wang, L., Li, G., and Fan, C. (2009) Goldnanoparticle-based multicolor nanobeacons for sequence-specific DNA analysis. Angewandte Chemie, International Edition, 48 (46), 8670–8674. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 262 473 Ray, P.C., Darbha, G.K., Ray, A., Walker, 474 475 476 477 478 479 480 481 J., and Hardy, W. (2007) Gold nanoparticle based FRET for DNA detection. Plasmonics, 2 (4), 173–183. Zhang, J., Wang, L., Zhang, H., Boey, F., Song, S., and Fan, C. (2010) Aptamerbased multicolor fluorescent gold nanoprobes for multiplex detection in homogeneous solution. Small, 6 (2), 201–204. Chen, Y., O’Donoghue, M.B., Huang, Y.F., Kang, H., Phillips, J.A., Chen, X., Estevez, M.C., Yang, C.J., and Tan, W. (2010) A surface energy transfer nanoruler for measuring binding site distances on live cell surfaces. Journal of the American Chemical Society, 132 (46), 16559–16570. Yeh, H.-Y., Yates, M.V., Mulchandania, A., and Chen, W. (2010) Molecular beacon– quantum dot–Au nanoparticle hybrid nanoprobes for visualizing virus replication in living cells. Chemical Communications, 46 (22), 3914–3916. Dyadyusha, L., Yin, H., Jaiswal, S., Brown, T., Baumberg, J.J., Booy, F.P., and Melvin, T. (2005) Quenching of CdSe quantum dot emission, a new approach for biosensing. Chemical Communications, (25), 3201–3203. Gueroui, Z. and Libchaber, A. (2004) Single-molecule measurements of goldquenched quantum dots. Physical Review Letters, 93 (16), 166108. Wargnier, R., Baranov, A.V., Maslov, V.G., Stsiapura, V., Artemyev, M., Pluot, M., Sukhanova, A., and Nabiev, I. (2004) Energy transfer in aqueous solutions of oppositely charged CdSe/ZnS core/shell quantum dots and in quantum dotnanogold assemblies. Nano Letters, 4 (3), 451–457. Peng, H.-I., Strohsahl, C.M., and Miller, B.L. (2012) Microfluidic nanoplasmonicenabled device for multiplex DNA detection. Lab on a Chip, 12 (6), 1089–1093. Xia, Y., Song, L., and Zhu, C. (2011) Turn-on and near-infrared fluorescent sensing for 2,4,6-trinitrotoluene based on hybrid (gold nanorod)-(quantum dots) assembly. Analytical Chemistry, 83 (4), 1401–1407. 482 Quach, A.D., Crivat, G., Tarr, M.A., and 483 484 485 486 487 488 489 490 491 Rosenzweig, Z. (2011) Gold nanoparticle– quantum dot–polystyrene microspheres as fluorescence resonance energy transfer probes for bioassays. Journal of the American Chemical Society, 133 (7), 2028–2030. Choi, S., Dickson, R.M., and Yu, J. (2012) Developing luminescent silver nanodots for biological applications. Chemical Society Reviews, 41 (5), 1867–1891. He, H., Xie, C., and Ren, J. (2008) Nonbleaching fluorescence of gold nanoparticles and its applications in cancer cell imaging. Analytical Chemistry, 80 (15), 5951–5957. Sonnichsen, C., Reinhard, B.M., Liphardt, J., and Alivisatos, A.P. (2005) A molecular ruler based on plasmon coupling of single gold and silver nanoparticles. Nature Biotechnology, 23 (6), 741–745. Tabor, C., Murali, R., Mahmoud, M., and El-Sayed, M.A. (2009) On the use of plasmonic nanoparticle pairs as a plasmon ruler: the dependence of the near-field dipole plasmon coupling on nanoparticle size and shape. The Journal of Physical Chemistry. A, 113 (10), 1946–1953. Liu, N., Hentschel, M., Weiss, T., Alivisatos, A.P., and Giessen, H. (2011) Three-dimensional plasmon rulers. Science, 332 (6036), 1407–1410. Liu, J., Lam, J.W.Y., and Tang, B.Z. (2009) Aggregation-induced emission of silole molecules and polymers: fundamental and applications. Journal of Inorganic and Organometallic Polymers and Materials, 19 (3), 249–285. Vaccaro, L., Morana, A., Radzig, V., and Cannas, M. (2011) Bright visible luminescence in silica nanoparticles. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 115 (40), 19476–19481. Belomoin, G., Therrien, J., and Nayfeh, M. (2000) Oxide and hydrogen capped ultrasmall blue luminescent Si nanoparticles. Applied Physics Letters, 77 (6), 779–781. Belomoin, G., Therrien, J., Smith, A., Rao, S., Twesten, R., Chaieb, S., Nayfeh, M.H., Wagner, L., and Mitas, L. (2002) j263 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 492 493 494 495 496 497 498 499 500 Observation of a magic discrete family of ultrabright Si nanoparticles. Applied Physics Letters, 80 (5), 841–843. Kapaklis, V., Politis, C., Poulopoulos, P., and Schweiss, P. (2005) Photoluminescence from silicon nanoparticles prepared from bulk amorphous silicon monoxide by the n. Applied Physics Letters, 87 (12), 123114. Prusty, S., Mavi, H.S., and Shukla, A.K. (2005) Optical nonlinearity in silicon nanoparticles: effect of size and probing intensity. Physical Review. B, Condensed Matter and Materials Physics, 71 (11), 113313. Kusova, K., Cibulka, O., Dohnalova, K., Pelant, I., Valenta, J., Fucikova, A., Zidek, K., Lang, J., Englich, J., Matejka, P., Stepanek, P., and Bakardjieva, S. (2010) Brightly luminescent organically capped silicon nanocrystals fabricated at room temperature and atmospheric pressure. ACS Nano, 4 (8), 4495–4504. Kang, Z., Liu, Y., and Lee, S.-T. (2011) Small-sized silicon nanoparticles: new nanolights and nanocatalysts. Nanoscale, 3 (3), 777–791. Wang, L., Reipa, V., and Blasic, J. (2004) Silicon nanoparticles as a luminescent label to DNA. Bioconjugate Chemistry, 15 (2), 409–412. Ramiro-Manzano, F., Fenollosa, R., XifrePerez, E., Garin, M., and Meseguer, F. (2012) Porous silicon microcavities: synthesis, characterization, and application to photonic barcode devices. Nanoscale Research Letters, 7, 497. Li, Z.F. and Ruckenstein, E. (2004) Watersoluble poly(acrylic acid) grafted luminescent silicon nanoparticles and their use as fluorescent biological staining labels. Nano Letters, 4 (8), 1463–1467. Park, J.-H., Gu, L., von Maltzahn, G., Ruoslahti, E., Bhatia, S.N., and Sailor, M.J. (2009) Biodegradable luminescent porous silicon nanoparticles for in vivo applications. Nature Materials, 8 (4), 331–336. Hong, C., Lee, J., Son, M., Hong, S.S., and Lee, C. (2011) In vivo cancer cell destruction using porous silicon nanoparticles. Anti-Cancer Drugs, 22 (10), 971–977. 501 Osminkina, L.A., Tamarov, K.P., Sviridov, 502 503 504 505 506 507 508 509 510 A.P., Galkin, R.A., Gongalsky, M.B., Solovyev, V.V., Kudryavtsev, A.A., and Timoshenko, V.Y. (2012) Photoluminescent biocompatible silicon nanoparticles for cancer theranostic applications. Journal of Biophotonics, 5 (7), 529–535. Singh, M.P., Atkins, T.M., Muthuswamy, E., Kamali, S., Tu, C., Louie, A.Y., and Kauzlarich, S.M. (2012) Development of iron-doped silicon nanoparticles as bimodal imaging agents. ACS Nano, 6 (6), 5596–5604. Li, X.G., He, Y.Q., Talukdar, S.S., and Swihart, M.T. (2003) Process for preparing macroscopic quantities of brightly photoluminescent silicon nanoparticles with emission spanning the visible spectrum. Langmuir, 19 (20), 8490–8496. Zou, J., Baldwin, R.K., Pettigrew, K.A., and Kauzlarich, S.M. (2004) Solution synthesis of ultrastable luminescent siloxane-coated silicon nanoparticles. Nano Letters, 4 (7), 1181–1186. Wang, J., Liu, Y., Peng, F., Chen, C., He, Y., Ma, H., Cao, L., and Sun, S. (2012) A general route to efficient functionalization of silicon quantum dots for highperformance fluorescent probes. Small, 8 (15), 2430–2435. Shen, P., Uesawa, N., Inasawa, S., and Yamaguchi, Y. (2010) Characterization of flowerlike silicon particles obtained from chemical etching: visible fluorescence and superhydrophobicity. Langmuir, 26 (16), 13522–13527. Roussel, M., Chen, W., Talbot, E., Larde, R., Cadel, E., Gourbilleau, F., Grandidier, B., Stievenard, D., and Pareige, P. (2011) Atomic scale investigation of silicon nanowires and nanoclusters. Nanoscale Research Letters, 6, 271. O’Farrell, N., Houlton, A., and Horrocks, B.R. (2006) Silicon nanoparticles: applications in cell biology and medicine. International Journal of Nanomedicine, 1 (4), 451–472. Sanderson, K. (2009) Quantum dots go large. Nature, 459 (7248), 760–761. Bruchez, M., Moronne, M., Gin, P., Weiss, S., and Alivisatos, A.P. (1998) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 264 511 512 513 514 515 516 517 518 519 520 521 Semiconductor nanocrystals as fluorescent biological labels. Science, 281 (5385), 2013–2016. Chan, W.C.W. and Nie, S.M. (1998) Quantum dot bioconjugates for ultrasensitive nonisotopic detection. Science, 281 (5385), 2016–2018. Medintz, I.L., Uyeda, H.T., Goldman, E. R., and Mattoussi, H. (2005) Quantum dot bioconjugates for imaging, labelling and sensing. Nature Materials, 4 (6), 435–446. Algar, W.R., Susumu, K., Delehanty, J.B., and Medintz, I.L. (2011) Semiconductor quantum dots in bioanalysis: crossing the valley of death. Analytical Chemistry, 83 (23), 8826–8837. Frasco, M.F. and Chaniotakis, N. (2010) Bioconjugated quantum dots as fluorescent probes for bioanalytical applications. Analytical and Bioanalytical Chemistry, 396 (1), 229–240. Esteve-Turrillas, F.A. and Abad-Fuentes, A. (2013) Applications of quantum dots as probes in immunosensing of small-sized analytes. Biosensors & Bioelectronics, 41, 12–29. Medintz, I.L. and Mattoussi, H. (2009) Quantum dot-based resonance energy transfer and its growing application in biology. Physical Chemistry Chemical Physics, 11 (1), 17–45. Zhang, Y. and Wang, T.-H. (2012) Quantum dot enabled molecular sensing and diagnostics. Theranostics, 2 (7), 631–654. Zhou, D. (2012) Quantum dot-nucleic acid/aptamer bioconjugate-based fluorimetric biosensors. Biochemical Society Transactions, 40, 635–639. Resch-Genger, U., Grabolle, M., CavaliereJaricot, S., Nitschke, R., and Nann, T. (2008) Quantum dots versus organic dyes as fluorescent labels. Nature Methods, 5 (9), 763–775. Pons, T., Medintz, I.L., Wang, X., English, D.S., and Mattoussi, H. (2006) Solutionphase single quantum dot fluorescence resonance energy transfer. Journal of the American Chemical Society, 128 (47), 15324–15331. Clapp, A.R., Pons, T., Medintz, I.L., Delehanty, J.B., Melinger, J.S., Tiefenbrunn, T., Dawson, P.E., Fisher, B. 522 523 524 525 526 527 528 529 R., O’Rourke, B., and Mattoussi, H. (2007) Two-photon excitation of quantum-dotbased fluorescence resonance energy transfer and its applications. Advanced Materials, 19 (15), 1921–1926. Rogach, A.L., Eychmueller, A., Hickey, S. G., and Kershaw, S.V. (2007) Infraredemitting colloidal nanocrystals: synthesis, assembly, spectroscopy, and applications. Small, 3 (4), 536–557. Moreels, I., Justo, Y., De Geyter, B., Haustraete, K., Martins, J.C., and Hens, Z. (2011) Size-tunable, bright, and stable PbS quantum dots: a surface chemistry study. ACS Nano, 5 (3), 2004–2012. Algar, W.R. and Krull, U.J. (2010) New opportunities in multiplexed optical bioanalyses using quantum dots and donor–acceptor interactions. Analytical and Bioanalytical Chemistry, 398 (6), 2439–2449. Jennings, T.L., Becker-Catania, S.G., Triulzi, R.C., Tao, G., Scott, B., Sapsford, K.E., Spindel, S., Oh, E., Jain, V., Delehanty, J.B., Prasuhn, D.E., Boeneman, K., Algar, W.R., and Medintz, I.L. (2011) Reactive semiconductor nanocrystals for chemoselective biolabeling and multiplexed analysis. ACS Nano, 5 (7), 5579–5593. Wu, C.-S., Oo, M.K.K., and Fan, X. (2010) Highly sensitive multiplexed heavy metal detection using quantum-dotlabeled DNAzymes. ACS Nano, 4 (10), 5897–5904. Suzuki, M., Husimi, Y., Komatsu, H., Suzuki, K., and Douglas, K.T. (2008) Quantum dot FRET biosensors that respond to pH, to proteolytic or nucleolytic cleavage, to DNA synthesis, or to a multiplexing combination. Journal of the American Chemical Society, 130 (17), 5720–5725. Kim, Y.-P., Oh, Y.-H., Oh, E., Ko, S., Han, M.-K., and Kim, H.-S. (2008) Energy transfer-based multiplexed assay of proteases by using gold nanoparticle and quantum dot conjugates on a surface. Analytical Chemistry, 80 (12), 4634–4641. Lowe, S.B., Dick, J.A.G., Cohen, B.E., and Stevens, M.M. (2012) Multiplex sensing of protease and kinase enzyme activity via orthogonal coupling of j265 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 530 531 532 533 534 535 536 537 quantum dot peptide conjugates. ACS Nano, 6 (1), 851–857. Zhang, Y. and Clapp, A. (2011) Overview of stabilizing ligands for biocompatible quantum dot nanocrystals. Sensors, 11 (12), 11036–11055. Muro, E., Pons, T., Lequeux, N., Fragola, A., Sanson, N., Lenkei, Z., and Dubertret, B. (2010) Small and stable sulfobetaine zwitterionic quantum dots for functional live-cell imaging. Journal of the American Chemical Society, 132 (13), 4556–4557. Susumu, K., Uyeda, H.T., Medintz, I.L., Pons, T., Delehanty, J.B., and Mattoussi, H. (2007) Enhancing the stability and biological functionalities of quantum dots via compact multifunctional ligands. Journal of the American Chemical Society, 129 (45), 13987–13996. Susumu, K., Oh, E., Delehanty, J.B., Blanco-Canosa, J.B., Johnson, B.J., Jain, V., Hervey, W.J., Algar, W.R., Boeneman, K., Dawson, P.E., and Medintz, I.L. (2011) Multifunctional compact zwitterionic ligands for preparing robust biocompatible semiconductor quantum dots and gold nanoparticles. Journal of the American Chemical Society, 133 (24), 9480– 9496. Halivni, S., Sitt, A., Hadar, I., and Banin, U. (2012) Effect of nanoparticle dimensionality on fluorescence resonance energy transfer in nanoparticle–dye conjugated systems. ACS Nano, 6 (3), 2758–2765. Artemyev, M., Ustinovich, E., and Nabiev, I. (2009) Efficiency of energy transfer from organic dye molecules to CdSe–ZnS nanocrystals: nanorods versus nanodots. Journal of the American Chemical Society, 131 (23), 8061–8065. Hardzei, M., Artemyev, M., Molinari, M., Troyon, M., Sukhanova, A., and Nabiev, I. (2012) Comparative efficiency of energy transfer from CdSe–ZnS quantum dots or nanorods to organic dye molecules. ChemPhysChem, 13 (1), 330–335. Zhang, L., Yin, L., Wang, C., Lun, N., Qi, Y., and Xiang, D. (2010) Origin of visible photoluminescence of ZnO quantum dots: defect-dependent and sizedependent. The Journal of Physical 538 539 540 541 542 543 544 545 Chemistry. C, Nanomaterials and Interfaces, 114 (21), 9651–9658. Obonyo, O., Fisher, E., Edwards, M., and Douroumis, D. (2010) Quantum dots synthesis and biological applications as imaging and drug delivery systems. Critical Reviews in Biotechnology, 30 (4), 283–301. Zhou, W., Schwartz, D.T., and Baneyx, F. (2010) Single-pot biofabrication of zinc sulfide immuno-quantum dots. Journal of the American Chemical Society, 132 (13), 4731–4738. Huang, H.-q., He, M.-x., Wang, W.-x., Liu, J.-l., Mi, C.-c., and Xu, S.-k. (2012) Biosynthesis of CdS quantum dots in saccharomyces cerevisiae and spectroscopic characterization. Spectroscopy and Spectral Analysis, 32 (4), 1090–1093. Mi, C., Wang, Y., Zhang, J., Huang, H., Xu, L., Wang, S., Fang, X., Fang, J., Mao, C., and Xu, S. (2011) Biosynthesis and characterization of CdS quantum dots in genetically engineered Escherichia coli. Journal of Biotechnology, 153 (3–4), 125–132. Cha, T.-G., Baker, B.A., Salgado, J., Bates, C.J., Chen, K.H., Chang, A.C., Akatay, M. C., Han, J.-H., Strano, M.S., and Choi, J. H. (2012) Understanding oligonucleotidetemplated nanocrystals: growth mechanisms and surface properties. ACS Nano, 6 (9), 8136–8143. Baker, B.A. and Choi, J.H. (2009) Oligonucleotide DNA and RNA as direct capping ligand for nanocrystals: an emerging method for biological diagnostics and therapeutics. Nano, 4 (4), 189–199. Medintz, I.L., Pons, T., Susumu, K., Boeneman, K., Dennis, A.M., Farrell, D., Deschamps, J.R., Melinger, J.S., Bao, G., and Mattoussi, H. (2009) Resonance energy transfer between luminescent quantum dots and diverse fluorescent protein acceptors. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 113 (43), 18552–18561. Dennis, A.M. and Bao, G. (2008) Quantum dot–fluorescent protein pairs as novel fluorescence resonance energy Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 266 546 547 548 549 550 551 552 553 transfer probes. Nano Letters, 8 (5), 1439– 1445. Mandal, G., Bardhan, M., and Ganguly, T. (2011) Occurrence of Forster resonance energy transfer between quantum dots and gold nanoparticles in the presence of a biomolecule. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 115 (43), 20840–20848. Haldar, K.K., Sen, T., and Patra, A. (2010) Metal conjugated semiconductor hybrid nanoparticle-based fluorescence resonance energy transfer. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 114 (11), 4869–4874. Buckhout-White, S., Ancona, M., Oh, E., Deschamps, J.R., Stewart, M.H., BlancoCanosa, J.B., Dawson, P.E., Goldman, E. R., and Medintz, I.L. (2012) Multimodal characterization of a linear DNA-based nanostructure. ACS Nano, 6 (2), 1026–1043. Hildebrandt, N. and Geissler, D. (2012) Semiconductor quantum dots as FRET acceptors for multiplexed diagnostics and molecular ruler application. Advances in Experimental Medicine and Biology, 733, 75–86. Medintz, I.L., Farrell, D., Susumu, K., Trammell, S.A., Deschamps, J.R., Brunel, F.M., Dawson, P.E., and Mattoussi, H. (2009) Multiplex charge-transfer interactions between quantum dots and peptide-bridged ruthenium complexes. Analytical Chemistry, 81 (12), 4831–4839. Medintz, I.L., Stewart, M.H., Trammell, S. A., Susumu, K., Delehanty, J.B., Mei, B.C., Melinger, J.S., Blanco-Canosa, J.B., Dawson, P.E., and Mattoussi, H. (2010) Quantum-dot/dopamine bioconjugates function as redox coupled assemblies for in vitro and intracellular pH sensing. Nature Materials, 9 (8), 676–684. Boulesbaa, A., Huang, Z., Wu, D., and Lian, T. (2010) Competition between energy and electron transfer from CdSe QDs to adsorbed rhodamine B. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 114 (2), 962–969. Morris-Cohen, A.J., Frederick, M.T., Cass, L.C., and Weiss, E.A. (2011) Simultaneous determination of the adsorption constant 554 555 556 557 558 559 560 561 and the photoinduced electron transfer rate for a CdS quantum dot–viologen complex. Journal of the American Chemical Society, 133 (26), 10146–10154. Medintz, I.L. and Deschamps, J.R. (2006) Maltose-binding protein: a versatile platform for prototyping biosensing. Current Opinion in Biotechnology, 17 (1), 17–27. Tavares, A.J., Noor, M.O., Vannoy, C.H., Algar, W.R., and Krull, U.J. (2012) Onchip transduction of nucleic acid hybridization using spatial profiles of immobilized quantum dots and fluorescence resonance energy transfer. Analytical Chemistry, 84 (1), 312–319. Prasuhn, D.E., Feltz, A., Blanco-Canosa, J. B., Susumu, K., Stewart, M.H., Mei, B.C., Yakovlev, A.V., Loukov, C., Mallet, J.-M., Oheim, M., Dawson, P.E., and Medintz, I. L. (2010) Quantum dot peptide biosensors for monitoring caspase 3 proteolysis and calcium ions. ACS Nano, 4 (9), 5487–5497. Medintz, I.L., Clapp, A.R., Brunel, F.M., Tiefenbrunn, T., Uyeda, H.T., Chang, E.L., Deschamps, J.R., Dawson, P.E., and Mattoussi, H. (2006) Proteolytic activity monitored by fluorescence resonance energy transfer through quantum-dotpeptide conjugates. Nature Materials, 5 (7), 581–589. Welser, K., Adsley, R., Moore, B.M., Chan, W.C., and Aylott, J.W. (2011) Protease sensing with nanoparticle based platforms. Analyst, 136 (1), 29–41. Dezhurov, S.V., Volkova, I.Y., and Wakstein, M.S. (2011) FRET-based biosensor for oleic acid in nanomolar range with quantum dots as an energy donor. Bioconjugate Chemistry, 22 (3), 338–345. Medintz, I.L., Berti, L., Pons, T., Grimes, A.F., English, D.S., Alessandrini, A., Facci, P., and Mattoussi, H. (2007) A reactive peptidic linker for selfassembling hybrid quantum dot–DNA bioconjugates. Nano Letters, 7 (6), 1741–1748. Zhang, C.-y. and Johnson, L.W. (2009) Single quantum-dot-based aptameric nanosensor for cocaine. Analytical Chemistry, 81 (8), 3051–3055. j267 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 6 Materials for FRET Analysis: Beyond Traditional Dye–Dye Combinations 562 Goldman, E.R., Medintz, I.L., Whitley, J. 563 564 565 566 567 L., Hayhurst, A., Clapp, A.R., Uyeda, H.T., Deschamps, J.R., Lassman, M.E., and Mattoussi, H. (2005) A hybrid quantum dot–antibody fragment fluorescence resonance energy transfer-based TNT sensor. Journal of the American Chemical Society, 127 (18), 6744–6751. Wei, Q., Lee, M., Yu, X., Lee, E.K., Seong, G.H., Choo, J., and Cho, Y.W. (2006) Development of an open sandwich fluoroimmunoassay based on fluorescence resonance energy transfer. Analytical Biochemistry, 358 (1), 31–37. Stringer, R.C., Schommer, S., Hoehn, D., and Grant, S.A. (2008) Development of an optical biosensor using gold nanoparticles and quantum dots for the detection of porcine reproductive and respiratory syndrome virus. Sensors and Actuators B: Chemical, 134 (2), 427–431. Geissler, D., Charbonniere, L.J., Ziessel, R.F., Butlin, N.G., Loehmannsroeben, H.G., and Hildebrandt, N. (2010) Quantum dot biosensors for ultrasensitive multiplexed diagnostics. Angewandte Chemie, International Edition, 49 (8), 1396–1401. Sener, M., Struempfer, J., Hsin, J., Chandler, D., Scheuring, S., Hunter, C.N., and Schulten, K. (2011) Forster energy transfer theory as reflected in the structures of photosynthetic lightharvesting systems. ChemPhysChem, 12 (3), 518–531. Stein, I.H., Steinhauer, C., and Tinnefeld, P. (2011) Single-molecule four-color FRET visualizes energy-transfer paths on DNA origami. Journal of the American Chemical Society, 133 (12), 4193–4195. 568 Boeneman, K., Prasuhn, D.E., Blanco- 569 570 571 572 573 Canosa, J.B., Dawson, P.E., Melinger, J.S., Ancona, M., Stewart, M.H., Susumu, K., Huston, A., and Medintz, I.L. (2010) Selfassembled quantum dot-sensitized multivalent DNA photonic wires. Journal of the American Chemical Society, 132 (51), 18177–18190. Kawahara, S., Uchimaru, T., and Murata, S. (1999) Efficiency enhancement of longrange energy transfer by sequential multistep FRET using fluorescence labeled DNA. Nucleic Acids Symposium Series, 42 (1), 241–242. Hippius, C., van Stokkum, I.H.M., Gsaenger, M., Groeneveld, M.M., Williams, R.M., and Wuerthner, F. (2008) Sequential FRET processes in calix-4-arene-linked orange-red-green perylene bisimide dye zigzag arrays. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 112 (7), 2476–2486. Tong, A.K. and Ju, J. (2006) Multiplex single-nucleotide polymorphism detection by combinatorial fluorescence energy transfer tags and molecular affinity. Methods in Molecular Biology (Clifton, N. J.), 335, 201–214. Smith, J.J., Conrad, D.W., Cuneo, M.J., and Hellinga, H.W. (2005) Orthogonal site-specific protein modification by engineering reversible thiol protection mechanisms. Protein Science, 14 (1), 64–73. Xiong, L., Shuhendler, A.J., and Rao, J. (2012) Self-luminescing BRET–FRET near-infrared dots for in vivo lymph-node mapping and tumour imaging. Nature Communications, 3, 1193. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 268 Part Two Common FRET Techniques/Applications FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j269 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine Samantha Spindel, Jessica Granek, and Kim E. Sapsford 7.1 Introduction F€orster (fluorescence) resonance energy transfer (FRET) is of particular interest to scientists due to its intrinsic sensitivity to small variations in molecular distance and orientation. When appropriately configured, qualitative and quantitative information about the system under study can be determined using FRET. FRET has numerous applications in the study of biological phenomena such as structure elucidation, inter- and intracellular processes, and in vivo diagnostics, which are covered in other chapters. FRET is increasingly being used in the area of molecular in vitro diagnostics (IVDs – which can be genomic, epigenetic, or protein-based assays) for sensing applications in the medical arena, such as clinical diagnostics/ prognostics and personalized medicine. The desire for high-throughput, real-time, simple, and rapid assays has been one of the driving forces behind the increased interest in FRET as a signal transduction mechanism. In particular, FRET has shown great potential in the area of point-of-care (PoC) diagnostics, which is discussed in more detail later [1]. Some inherent complications of FRET using more traditional reagents include photobleaching, spectral cross talk, and direct excitation of the acceptor, which can all increase background and reduce sensitivity. The production of high-quality reagents that address some of these issues, along with substantial progress in the field of nanotechnology, have increased the utility of FRET [2]. There are a number of potential in vitro target analytes of interest, including proteins, metabolites, and other small molecules, such as drugs (both recreational and doping), toxins, nucleic acids, human cells, microbes, and other pathogens highlighted throughout this chapter [1]. Target analytes that fall under the label of “biomarker” are of particular interest from a clinical diagnostic and personalized medicine perspective [3]. The consensus definition of a biomarker is “a characteristic that is objectively measured and evaluated as an indicator of normal biologic processes, pathogenic processes or pharmacologic responses to a therapeutic intervention” [4]. Biomarkers have a range of potential clinical utilities, including identifying the presence of disease and characterizing disease subtype FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j271 j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine (diagnosis), determining disease prognosis, aiding in selection of appropriate therapeutic doses, predicting clinical benefit or adverse response to therapy, and monitoring treatment outcomes [3]. FRET configurations used for sensing these target analytes are diverse in nature, relying on a number of potential mechanisms, including cleavage, binding, sensitive absorption/emission profiles, and/ or structural rearrangements (conformational changes), which ultimately alter the FRET between the donor/acceptor molecule(s), resulting in a measurable change in the FRET signal. Examples of these various mechanisms will be highlighted throughout this chapter. We have chosen to discuss FRET applications in terms of the “sensing/recognition” molecules that interact with the target analyte of interest, resulting in measurable FRET changes. These sensing/recognition molecules can take a variety of forms, including small organic molecules, polymers, carbohydrates, antibodies, proteins and peptides, nucleic acids [including deoxyribonucleic acid (DNA) and ribonucleic acid (RNA)], and aptamers. 7.2 Small Organic Molecules and Synthetic Organic Polymers There are a number of organic molecules (sometimes referred to as probes) that have unique absorption or emission profile responses to particular analytes, which make them useful for FRET configurations [5]. The group of Lakowicz as well as others, for example, has developed a number of lifetime-based FRET sensors based upon organic molecules sensitive to changes in pH, NH3, and/or CO2 [6–9]. Although the spectroscopic change (absorbance or fluorescence-based) of the organic molecule is often measured directly [10], when the organic molecule has an absorbance-based response to an analyte, incorporating a fluorescence component to the detection mechanism, such as via FRET may improve assay sensitivity [7,9]. Ruedas-Rama and Hall developed a FRET-based sensor that measures pH changes in response to the activity of enzymes such as urease and creatinine deiminase by immobilizing the enzymes and a pH-sensitive calcium red dye (CaR) to the surface of quantum dots (QDs) [11]. The established interaction between boronic acid moieties and various cis-1,2- or 1,3-diol-containing molecules, such as saccharides and dopamine, has led some researchers to develop FRET-based glucose or sugar sensors [12,13]. For example, Freeman and coworkers created fluorescently labeled galactose, glucose, and dopamine as analyte analogues in competitive assays that used phenyl boronic acidfunctionalized QDs as the FRET donor. In the presence of the analyte (galactose, glucose, or dopamine), the labeled sugar/dopamine (analyte analogue) was displaced from the QD, reducing FRET and increasing QD emission. Wang et al. designed thermoresponsive poly(N-isopropylacrylamide) (PNIPAM) microgels that incorporated FRET donor and acceptor dyes along with N-acryloyl-3-aminophenylboronic acid (APBA) for glucose and temperature sensing [13]. Increased temperature caused the microgels to shrink, enhancing the FRET signal, while the addition of glucose under specific temperature conditions caused the microgels to expand, Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 272 weakening the FRET signal. A similar concept for Kþ sensing was achieved by replacing the boronic acid moieties with crown ethers [14]. The response of various polymeric materials to pH or protein adsorption has been investigated using FRET transduction [15–18]. Chiu et al. used the pHsensitive conformational change that occurs in N-palmitoyl chitosan (NPCS), conjugated with a Cy3 [Cyanine 3 (dye)]donor dye and a Cy5 [Cyanine 5 (dye)] acceptor dye, to measure pH changes in biological environments via FRET (Figure 7.1) [16]. 7.3 Carbohydrate–Lipid Due to the numerous physiological roles carbohydrates play in human biology, there has been increasing interest in their use as recognition elements for sensing applications, especially for protein and bacterial targets, although their combination with FRET is presently somewhat limited [19,20]. Song et al. designed a multi-FRET system using labeled GM1 (monosialotetrahexosylgangliosides) (carbohydrate/lipid components of the cell plasma membrane) to monitor their interaction with cholera toxin [21]. Ma and Cheng developed a mix-and-detect assay using dye-labeled GM1 and polydiacetylene (PDA) vesicles [22]. Cholera toxin bound the dye-labeled GM1, preventing GM1 incorporation into the PDA vesicles and its subsequent quenching via FRET. McGiven et al. developed a competitive assay for detection of anti-Brucella antibodies in serum using Brucella smooth lipopolysaccharide (LPS –carbohydrate/ lipid structure found in the outer membrane of Gram-negative bacteria) and antiBrucella monoclonal antibody labeled with a long-lifetime donor terbium chelate fluorophore [23]. The use of time-resolved fluorescence allowed a significant reduction in the background fluorescence of the serum sample matrix and therefore increased the sensitivity of anti-Brucella detection. 7.4 The Biotin–Avidin Interaction The avidin–biotin complex is worth highlighting, as it represents one of the strongest known noncovalent interactions (Ka 1015 M1) and as such it is often used as a model system for protein–ligand interactions and for testing new materials for sensing applications [24–27]. The latter is particularly true in FRET studies, where researchers have used the avidin–biotin interaction to test new donor– acceptor combinations, including luminescent lanthanide complexes (LLCs) and nanoparticles (NPs) (europium and terbium) with QDs [28–30], gold nanoparticles (Au NPs) with QDs [31], conjugated polyelectrolytes and tetramethylrhodamine (TMR) [32], dye-labeled silica NPs with various dyes [33], and a two-photon excitable organic molecule with a dark quencher [34]. While the detection of biotin is the more clinically relevant of the biotin–avidin pair, since it is an essential component of j273 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.4 The Biotin–Avidin Interaction j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine Figure 7.1 FRET-based pH-responsive polymer sensors. (a) Design and working principle of the FRET-based pH-responsive polymer sensor developed by Chiu et al. [16]. N-palmitoyl chitosan (NPCS) was labeled with either Cy3 (donor) or Cy5 (acceptor) before being mixed to form the FRET sensor. Changes in pH caused a conformational change in the polymer structure that resulted in measurable changes in the FRET signal. (b) Nonnormalized FRET spectra showing the increase in FRET obtained from the nanosensors as the pH decreases from 8.0 to 4.0. (c) Dual emission images of Cy3/Cy5/chitosan – 15% nanoparticle suspensions obtained with an in vivo imaging system. (Reprinted with permission from Ref. [16]. Copyright 2010, American Chemical Society.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 274 vitamin B, most researchers use this complex for avidin sensing in a competitive assay format [31,32,34]. 7.5 Proteins and Peptides There is a wide range of protein- and peptide-based recognition events that can be probed using FRET transduction and hence used for sensing applications. Antibodies are an obvious example and are discussed separately in Section 7.6. The main examples here include binding proteins, epitope-derived peptide sequences, and protease-specific peptide sequences. 7.5.1 Binding Proteins There are a wide variety of binding proteins, in addition to antibodies, with sites that recognize and bind specific analytes, which make them attractive elements for FRET-based sensing. A popular example of binding proteins is the protein switch, which is composed of proteins that undergo conformational changes upon either covalent modifications or molecular recognition [35]. The conformational changes of protein switches such as periplasmic binding proteins (PBP) have routinely been studied via FRET [35–37], with glucose binding protein (GBP) and maltose binding protein (MBP), of particular interest due to the clinical relevance of their target analytes [38–39]. These hinge-based proteins typically transition from an open to a closed conformation upon target binding, somewhat akin to the Venus flytrap. A number of researchers have used this inherent conformational change for FRET-based glucose and maltose sensing [38,40–46]. Two main issues with GBP sensors are that they possess a narrow dynamic range and are oftentimes too sensitive for clinical use, detecting glucose in the micromolar range as opposed to the more desirable millimolar range. Jin et al. attempted to address this issue by developing a series of GBP mutants, one of which demonstrated a more physiologically relevant glucose detection range of 0–32 mM [46]. By using the conformation change that occurs in bovine serum albumin (BSA) when it binds to long chain fatty acids, Dezhurov et al. developed a BSA-based sensor for oleic acid. Here, BSA labeled with an acceptor dye was self-assembled onto CdSe/ ZnS QDs (donor), resulting in FRET. The binding of oleic acid to BSA caused a conformational change in BSA, decreasing the distance between the donor and acceptor, resulting in an increase in FRET efficiency [47]. In an alternative approach that relies on conformational change, Thurley et al. designed hairpin peptide beacons, analogous to DNA-based molecular beacons (MBs), comprising a protein-specific peptide sequence (that recognizes the target protein) flanked by two complementary peptide nucleic acid (PNA) arm segments, labeled with a donor– acceptor pair [48]. Binding of the target protein, in this case Src kinase, resulted in a j275 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.5 Proteins and Peptides j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine conformational change within the hairpin peptide beacon that ultimately separated the PNA arms and hence changed the FRET signal. As an alternative approach to using a protein’s inherent conformational change (which can produce a very tight sensor dynamic range) or for binding proteins that do not undergo significant structural changes upon target analyte binding, researchers have investigated the competitive displacement format. Medintz et al. demonstrated a number of iterations of this assay format using MBP for maltose detection, taking advantage of the weaker association of MBP for b-cyclodextrin than its target analyte maltose (Figure 7.2) [41,49–51]. Labeled b-cyclodextrin was allowed to bind to labeled MBP, resulting in FRET. Upon addition of maltose, the b-cyclodextrin was displaced (through competitive displacement) and a change in FRET resulted. Initial studies investigated the use of standard dye–dye or dye–quencher combinations for Figure 7.2 Competitive displacement assays using binding proteins. (a) Schematic representation of homogeneous FRET-based competitive displacement assay. Here, the donor-labeled binding protein interacts with an acceptor-labeled target analogue resulting in FRET. Introduction of the target analyte causes displacement of the target analogue, disrupting the FRET and causing an increase in the donor emission. (Adapted with permission from Ref. [49].) (b) Schematic representation of heterogeneous FRET-based competitive displacement assay. Here, the acceptor-labeled (in this case quencher) binding protein interacts with a donor-labeled target analogue, resulting in FRET, with both components immobilized on a microplate well. Introduction of the target analyte causes displacement of the target analogue, disrupting the FRET and causing an increase in the donor emission. Due to the immobilization of the sensor components, the system can be regenerated by simply washing away the target analyte. (Adapted with permission from Ref. [51].) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 276 the donor–acceptor FRET pair [49], while later studies focused on the use of QD donors for the design of single- and multi-FRET systems [50]. A unique system capable of regeneration, a highly desired biosensor quality, was also designed to overcome the single-use limitation (Figure 7.2b) [51]. Here, biotin-labeled MBP was bound to a NeutrAvidin coated surface and b-cyclodextrin was immobilized in close proximity to the MBP, via a flexible biotinylated DNA oligonucleotide. Varying the MBP mutant and DNA linker arm attachment allowed the ability to tune the sensor’s sensitivity and dynamic range, significantly improving the utility of the configuration. Competitive displacement assays have also been adopted for detection of glucose with glucokinase or concanavalin A (ConA) acting as the binding protein [52–59]. Cheung et al. developed a glucose sensor based on TMR-ConA and fluorescein-dextran (Mw 200 kDa) encapsulated within hydrogel pads and subsequently overcoated with polyelectrolyte multilayers, via layer-by-layer self-assembly [58]. The resulting sensor could measure glucose from 0–10 mM, had selective permeability – only allowing monosaccarides and disaccarides to reach the sensor biochemistry – and was reusable. Kim et al. used QD-labeled ConA in combination with Au NP-labeled dextran to sense protein glycosylation (specifically mannosylated proteins), which have been implicated in a number of diseases [60,61]. There are also a few instances in the literature where researchers have used the target binding or environmental spectral sensitivity of a protein to develop FRETbased sensors [62–64]. Stianese et al., for example, used the absorbance spectral changes that occur in cytochrome c peroxidase (CcP) upon binding nitric oxide (NO) to develop a FRET-based NO sensor by modifying the CcP with a near-infrared (IR) fluorescent dye [63]. Dennis et al. used the pH environmentally sensitive fluorescent protein (both excitation and emission spectra varied with pH), mOrange, coupled with QDs to develop a ratiometric pH sensor [64]. 7.5.2 Antigens and Epitope-Based Peptide Sequences Detection of antibodies is important from a clinical perspective as they can be used to diagnose exposure to infectious agents, autoimmune diseases, and allergies [65]. A couple of interesting FRET sensor designs have been proposed for antibody detection and involve using either the full antigen (the antibody’s target analyte) or epitopebased peptide sequences derived from the antibody’s target antigen (Figure 7.3). Sukhanova et al. measured autoantibodies of systemic sclerosis using QD-coded microbeads functionalized with the antigen topoisomerase I (topoI) in a flow cytometry-type assay [66]. Once the topoI autoantibodies in the patient sera bound the microbeads, dye-labeled antibodies then bound to the microbead captured autoantibodies, resulting in FRET. Rather than using the full antigen, Tian and Heyduk designed a FRET assay using two epitope-based peptide sequences, derived from cardiac troponin I, which is used to diagnose acute myocardial infarction (AMI) (Figure 7.3a) [67]. The two epitope-based peptide sequences were conjugated to flexible linkers modified with dye-labeled complementary oligonucleotides. Introduction of the antibody to cardiac troponin I caused the peptide sequences to bind to the j277 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.5 Proteins and Peptides j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine Figure 7.3 Antibody detection using epitope peptide FRET-based sensors. (a) Schematic representation of a FRET-based sensor using two separate antibody epitopes, each conjugated with short complementary oligonucleotides via flexible linkers and labeled with fluorescent probes. The oligonucleotides were designed to be short enough to not hybridize is the absence of the target antibody. Hybridization occurs in the presence of the target antibody, resulting in efficient FRET between the donor/acceptor fluorophores. (Adapted with permission from Ref. [67].) (b) Schematic representation of a FRET-based sensor using two antibody epitopes linked via a flexible linker and labeled with fluorescent probes. The fluorescent probes in this case were cerulean and Citrine fluorescent proteins designed with hydrophobic surface mutations that caused the proteins to interact in the absence of the target antibody, resulting in efficient FRET. Introduction of the target antibody resulted in separation of the fluorescent proteins measured as a decrease in FRET efficiency. (Adapted with permission from Ref. [69].) antibody, bringing the complementary oligonucleotides into close proximity, allowing them to hybridize and cause an increase in the FRET signal. To improve the multiplexing ability of this assay, the technology has been transitioned to a solid-surfacebased format for the detection of a range of target analytes, including antibodies and bacteria [68]. Golynskiy et al. developed a FRET based sensor for human immunodeficiency virus (HIV)-1 antibody in serum and saliva [69]. High FRET is observed in the absence of the antibody, where hydrophobic mutations on the protein surface caused a closed conformation of the fluorescent domains of the FRET-based sensor. In the presence of the antibody, bivalent binding between the antibody and two antigen epitope regions on the FRET-based sensor caused separation of fluorescent domains and a decrease in FRET (Figure 7.3b). Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 278 7.5.3 Peptide Sequences for Enzymatic Sensing Proteolytic enzymes (proteases) selectively cleave peptide bonds and have important physiological roles in many biological processes [70,71]. From a clinical perspective, they are key participants that are up- or downregulated in a number of diseases, and as such a variety of peptide-based assays have been developed to detect and monitor protease activity [70–73]. Many of these assays take advantage of FRET for detection; the basic sensing scheme is shown in Figure 7.4a, with some examples summarized in Table 7.1 and highlighted later. In general, a peptide sequence containing the protease-specific substrate is labeled with the donor and acceptor species, flanking the cleavage site. The close proximity of the donor–acceptor pair on the intact peptide sequence results in FRET. Protease activity causes peptide cleavage, increasing the distance between the donor/acceptor molecules and hence decreasing FRET, which is typically measured as increasing donor emission intensity. Proteases have a range of substrate specificities and many require only limited sequence recognition prior to cleavage. For example, trypsin recognizes single amino acids, cleaving the carboxyl side of lysine (K) and arginine (R) residues (except when followed by proline-P), while caspase-3 is tetrapeptide specific, recognizing the sequence motif D-X-X-D and cleaving after the aspartic acid residues (D) (X represents any amino acid). In contrast, botulinum neurotoxin A, which comprises a light chain protease and a heavy chain receptor binding protein, binds an optimum peptide substrate sequence of 16 residues, found in its physiological target synaptic protein SNAP25, prior to cleavage [81,82]. The specificity of the protease and the designed FRET substrate can be an issue when these types of sensors are applied to complex clinical samples, such as blood or sera, which may contain a number of physiological proteases that might increase the risk of false-positive responses. In some instances, purification, such as immunoextraction, or inclusion of inhibitors, which inhibit proteases other than the desired target, may be required to improve specificity. The serine protease trypsin, found in the digestive system, is a common target enzyme used by researchers when demonstrating initial proof-of-concept FRET protease assays, in part due to its well-characterized nature, single amino acid specificity (K and R residues), and commercial availability [83–87]. Trypsin is particularly popular with researchers when demonstrating new FRET donor–acceptor combinations or materials, such as silica nanobeads or QDs [83–86], or new detection platform technologies [86]. The group of Medintz and coworkers has developed a number of QD-peptide-FRET-based sensing platforms for a range of clinically relevant proteases, including caspase-3 (found to be downregulated in a number of cancers) [75,76], caspase-1 (a mediator for inflammation), thrombin (an important blood clotting protein), collagenase (an enzyme involved in cancer metastasis) [78], and botulinum neurotoxin A [79]. Cathepsin S, a cysteine protease that has been associated with obesity, atherosclerosis, and Alzheimer’s disease, was measured using an Abz-LEQ-EDDnp FRET peptide substrate [88]. Once optimized in buffer, the assay was applied to tissue homogenates, where a cocktail of inhibitors was required to prevent hydrolysis of the FRET peptide substrate by other j279 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.5 Proteins and Peptides j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine Figure 7.4 Protease sensing using peptide FRET-based sensors. (a) Schematic representation of a FRET-based sensor comprising a relatively short peptide sequence that encompasses the specific protease cleavage site flanked on either side by fluorescent probes. When intact, the fluorescent probes are in close proximity and undergo efficient FRET. However, introduction of the protease causes peptide cleavage, resulting in disruption of FRET and an increase in donor emission. (b) Factor Xa detection using a unique sequential BRET–FRET combination, termed SRET. The peptide-based probe comprised a peptide sequence (containing the protease-specific cleavage site) flanked by a thermostable firefly luciferase that produces yellow-green bioluminescence, and a red fluorescent protein labeled with a near-infrared fluorescent dye (AlexaFluor 680). When intact, the peptide probe undergoes efficient BRET/ FRET resulting in acceptor emission of the nearinfrared fluorescent dye (at 705 nm). Addition of the protease factor Xa results in a decrease in acceptor emission, due to disruption of the BRET/FRET process, as illustrated in the time course spectra. (Reprinted with permission from Ref. [91]. Copyright, Elsevier.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 280 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine physiological proteases such as the serine, metallo, aspartic, calcium-dependent, and caspase families. A FRETpeptide cleavage mechanism was also proposed for the potential diagnosis of thrombotic thrombocytopenic purpura (TTP) by measuring the decrease in ADAMTS13 activity [89,90]. Brancini et al. took the FRET peptide cleavage format a step further, designing a sequential bioluminescence resonance energy transfer (BRET)–FRET multienergy transfer assay platform based on firefly luciferase bioluminescence, for the detection of caspase-3, thrombin, and factor Xa activity [91]. The sensor was composed of firefly luciferase, which generates yellow-green bioluminescence and a red fluorescent protein (RFP) covalently labeled with a near-infrared fluorescent dye (Figure 7.4b). The firefly luciferase and RFP are connected via a decapeptide sequence, containing the protease-specific recognition/binding site; BRET occurs upon the addition of luciferase substrates. The subsequent FRET occurs between the RFP and the near-infrared fluorescent label. The advantage of this system is that BRET is activated via addition of chemical substrates rather than an external excitation source, making the signal background extremely low. Also, spectral resolution between the firefly luciferase bioluminescence (560 nm) and the nearinfrared fluorescent label (Em 705 nm) is excellent due to the sequential BRET– FRET format, improving assay sensitivity. BRET alone has also been demonstrated for measuring the protease activity of thrombin with improved sensitivity versus FRET [92]. Using a slightly modified format to that described in Figure 7.4a, Gratz et al. developed a FRET peptide assay for measuring the activity of the protein kinase CK2, which has been found to be upregulated in a number of cancers [93]. The assay relies on the ability of CK2 to block proteolytic cleavage of a FRET peptide substrate by CK2 serine phosphorylation of the elastase substrate S–D cleavage site. In addition to physiological human proteases, the detection of specific bacterial or viral enzymes/proteases can be used to indirectly diagnose infection, such as the bacteria Gram-positive Bacillus anthracis, the causative agent of anthrax, and the Clostridium species botulinum and tetani, as well as the virus severe acute respiratory syndrome coronavirus (SARS-CoV) [79,94–99]. Such rapid assays also provide the ability to screen for potential inhibitors that may prevent bacterial disease progression [95]. 7.6 Antibodies The ability to generate antibodies to a wide range of target analytes and their resulting specific nature make them very attractive biorecognition elements for FRET-based biosensing and diagnostics [100]. The sandwich assay represents the mainstream immunoassay format; however, it can be less practical from a FRET perspective due to the relatively large dimensions of the resulting antibody–antigen– antibody complex [20]. Researchers have overcome this limitation by using newer FRET materials, such as QDs or lanthanide complexes, unique labeling strategies, or through the use of antibody fragments, which for various reasons can improve FRET Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 282 efficiency, as illustrated in Figures 7.5–7.7 [101–109]. Wei et al., for example, took advantage of the high quantum yields of QDs to extend the F€ orster distance R0, measuring estrogen receptor b-antigen using full antibodies in a sandwich format [101] (Figure 7.5). The group of Hildebrandt and coworkers have developed homogeneous sandwich immunoassays using QD/lanthanide FRET combinations for detection of prostate-specific antigen [108] and alpha-fetoprotein (AFP) [109] (Figure 7.5b). Use of the long-lifetime lanthanide complexes allowed time-resolved measurements, which decreased background and improved assay sensitivity. Heyduk and coworkers devised a unique oligonucleotide antibody-labeling system that has been used by others to detect a range of target analyte types (both protein and bacterial) via a homogeneous sandwich assay [107,110,111]. The methodology uses antibodies, specific for distinct epitopes on the target analyte, labeled with short complementary oligonucleotides, and modified with either a donor or an acceptor species. Binding to the target analyte brings the complementary single-stranded deoxyribonucleic acid (ssDNA) into close proximity, resulting in hybridization and hence FRET (Figure 7.6). They have also modified the technology for use with solidsurface-based assays to facilitate multiplexing [68]. Chemiluminescence resonance energy transfer (CRET) between luminol and various acceptors (graphene nanosheets and Au NPs), catalyzed by horseradish peroxidase (HRP), has been demonstrated for detection of C-reactive protein, a biomarker of inflammation and cardiovascular diseases [106] and alpha-fetoprotein, a cancer biomarker [112]. An alternative approach to improving FRET efficiency for sandwich formats is to decrease the overall distance between the donor and the acceptor dyes through the use of antibody fragments (Figure 7.7). Sasajima et al. used fragments from the antibody variable domain (Fv) to develop an immunoassay for tyrosine phosphorylation [103], while Ohiro et al. used the larger Fab (fragment, antigen binding) antibody region combined with a leucine zipper motif for human serum albumin detection [102]. Other suitable antibody formats that have proven successful for target detection include competitive, competitive displacement, and direct detection, as illustrated in Figure 7.8 [20]. Competitive and competitive displacement assays are popular, although they require the design of a labeled target analyte analogue [20,67,100,113–116]. For example, Tan et al. demonstrated a competitive assay for D9-tetrahydrocannabinol, a major component of cannabis, in saliva (Figure 7.8a) [114], while Kattke et al. used a competitive displacement assay to detect the presence of a mold Aspergillus amstelodami [116]. Here, quencher-labeled Aspergillus fumigatus was incubated with anti-Aspergillus antibody conjugated with QDs, with the resulting antibody–antigen complex causing a significant reduction in QD emission due to FRET. Addition of the target analyte, A. amstelodami (for which the antibody had a higher affinity), caused displacement of the quencher-labeled A. fumigatus and an increasing QD emission. An alternative direct detection technique was developed by the Grant group based on the conformational change an antibody undergoes upon antigen binding (Figure 7.8b) [117]. This format has been used to detect the pathogens Listeria [118], Salmonella [118,119], porcine reproductive and respiratory syndrome virus (PRRSV) [120], and the clinical analytes cardiac troponin Tand I [121]. In the case j283 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.6 Antibodies j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine Figure 7.5 FRET-based sandwich immunoassays. (a) Schematic representation of a FRET-based sandwich immunoassay. Two antibodies (blue and purple) specific for different epitopes on the target analyte are labeled with either the donor or the acceptor fluorescent probe. Introduction of the target analyte brings the antibodies into close proximity, resulting in FRET. (b) Examples of some of the unique materials used to improve the FRET efficiency, when designing a sandwich immunoassay. Here, antibodies specific to alpha-fetoprotein (AFP) are labeled with either QD-doped microparticles or luminescent terbium chelates (LTCs). Introduction of AFP brings the antibodies into close proximity, resulting in FRET between the LTC donor and QD acceptor species. The emission (LTC and QD) along with the absorption (QD) spectra are illustrated along with the timeresolved luminescent decay measurements of the QDs alone (red squares), LTC alone (black triangles), and the QD– antibody1–AFP–antibody2–LTC complex (blue circles). (Reprinted with permission from Ref. [109]. Copyright, Elsevier.) Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 284 Figure 7.6 FRET-based sandwich immunoassays using unique oligonucleotide FRET-based labels. (a) Schematic representation of the FRET-based sandwich immunoassay. Two antibodies (blue and purple) specific for different epitopes on the target analyte are labeled with either the donor or the acceptor fluorescent probe, comprising short complementary oligonucleotides labeled with fluorescent probes and connected to the antibody via flexible linkers. Hybridization occurs in the presence of the target analyte, resulting in efficient FRET between the donor/ acceptor fluorophores. (Adapted with permission from Refs [107,110].) (b) Example of the above FRET-based sandwich immunoassay used for the detection of E. coli O157:H7 cells. Fluorescent images of the 96well microplate wells containing indicated amounts of target E. coli O157:H7 cells or control E. coli K12 cells. Upper panel shows donor emission image (Ex 488 nm and Em 530 nm) and lower panel shows acceptor emission due to FRET (Ex 488 nm and Em 690 nm). (Reprinted with permission from Ref. [110]. Copyright, Elsevier.) j285 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.6 Antibodies j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine Figure 7.7 FRET-based sandwich immunoassays using antibody fragments. (a) Schematic of a full antibody, highlighting the antibody fragments commonly incorporated into FRET-based assays. Both the Fv fragments – VH (variable heavy)/VL (variable light), and the Fab fragments contain the antigen-binding site of the antibody (illustrated as the red region on the antibody illustration). (b) FRET-based sandwich immunoassays using Fv antibody fragments. Fv fragments from the same antibody, comprising the VH and VL regions are labeled with donor or acceptor species. Introduction of the target antigen brings these Fv fragments into close proximity, resulting in efficient FRET. (c) FRET-based sandwich immunoassays using Fab antibody fragments. Fab fragments from two antibodies (blue and purple) specific for different epitopes on the target antigen are labeled with donor or acceptor species. Introduction of the target antigen brings these Fab fragments into close proximity, resulting in efficient FRET. of PRRSV detection, the carrier protein (Protein A) was conjugated to either an Au NP or a QD before the fluorescently labeled antibody was introduced, resulting in the FRET-sensing complex [120]. As illustrated in Figure 7.8b, antigen binding – in this case PRRSV – resulted in a conformational change in the antibody structure, resulting in a measurable change in the FRET signal. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 286 Figure 7.8 Alternative FRET-based immunoassay formats. (a) Competitive displacement immunoassay format. Donorlabeled antibody is preincubated with acceptorlabeled antigen analogue, resulting in FRET. Introduction of the target antigen causes displacement of the antigen analogue, resulting in a decrease in FRET and an increase in donor emission. Note that if the antigen analogue and target antigen are simultaneously incubated with the antibody, the format is referred to as a competitive format. (b) Direct immunoassay format. Here, the acceptor-labeled antibody is modified with a donor-labeled carrier molecule (CM), typically protein A or G, resulting in FRET. Introduction of the target antigen causes a conformational change in the antibody structure, altering the distance between the donor–acceptor pair and therefore changing the measured FRET signal. 7.7 Nucleic Acid (DNA/RNA) One of the main areas in which FRET has played a significant role in clinical applications is in nucleic acid-based genetic testing technologies. Completion of the Human Genome Project, concomitant whole genome sequencing of model organisms, the advent of next generation sequencing technologies, and genome-wide association studies have revolutionized our understanding of the role of genetic variation in human disease [122–125]. There are a number of genetic variations found in the human genome, including single nucleotide variants [single nucleotide polymorphisms (SNPs), insertions, or deletions – indels] and structural variants j287 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.7 Nucleic Acid (DNA/RNA) j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine [copy number variations (CNVs), insertions, deletions, inversions, repeats, translocations, duplications, etc.] [123,126]. SNPs (substitutions) represent the most frequently studied genetic variation to date [127]. Through genome-wide association studies, SNPs have been found to be important markers linking sequence variation to phenotypic changes for a number of diseases [123,127,128]. While FRET can be applied to DNA sequencing in the form of FRET primers, FRET terminators, and FRETcassettes [20,122,124,129], FRET has had a much more significant impact on molecular diagnostic technologies for detecting specific, relatively short DNA or RNA sequences (typically identified during DNA sequencing) for infectious disease diagnosis, genotyping, and pharmacogenomics [127,130– 136]. These technologies rely on the detection of target DNA hybridization and/or amplification. Hybridization and FRET have been applied to DNA microarrays for DNA and RNA detection [130,131,137–141]. However, real-time, homogeneous, and high-throughput assays are desired and FRET-based technologies, such as molecular beacons and amplification primers or probes, can uniquely address these requirements [20,130,133,135]. Some of the more common technologies are summarized later followed by a discussion of their medical application. 7.7.1 Molecular Beacons Since their inception in 1996, MBs [142] have been employed to detect DNA hybridization and select DNA sequences, gene mutations (e.g., SNPs), proteins, viruses, and changes in mRNA [130,139,143,144]. MBs have been used for a variety of other purposes as well such as to monitor polymerase chain reaction (PCR) amplification in real time. MBs are single-stranded nucleic acid probes that in their native state adopt a stem–loop or hairpin structure (Figure 7.9a). They are typically 25–35 nucleotides in length, comprising three main parts: (i) a loop portion (15–30 nucleotides) complementary to the known target DNA, (ii) a stem portion (5–8 nucleotides) consisting of two complementary arm sequences annealed on either side of the target sequence, and (iii) FRET donor/acceptor molecules labeled at the 50 and 30 ends [20,139,144,145]. In the presence of a target sequence, the molecular beacon unravels, separating the donor and acceptor, resulting in a measurable change in the FRET signal. A major advantage of MBs is that the target DNA does not need to be labeled, and multiplexing can be accomplished by selecting different donors and/or acceptors [130,146,147] or applying wavelength-shifting molecular beacons [143,145]. However, as with any of these assays, optimization is a key step to selective and sensitive detection of the target, and MB design is a fundamental aspect for the successful application of these types of probes. The distance between the FRETpair on the MB is important, and ideally the probe sequence should be more than double the length of the stem portion to ensure the FRET pair are far enough apart after hybridization to generate a significant FRET response. The melting temperature of MBs depends on the control of pH and temperature as well as length of the stem, the GC (guanine/ cytosine) content, and the ionic concentration of the buffer [144]. Although MBs are somewhat limited in their use because they can only detect relatively short DNA Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 288 Figure 7.9 Hybridization probes for nucleic acid-based assays. (a) Molecular beacons (MBs) comprise a stem–loop DNA structure that unravels in the presence of the target DNA, resulting in a decrease in FRET. (b) Probe–probe-based FRET hybridization probes comprise two probes (labeled with either donor or acceptor species) that upon hybridization to the target DNA come into close proximity, resulting in FRET. A subsequent melting curve analysis can be used to locate SNPs. sequences [144,146], once optimized, they can be used as detection probes, following amplification (typically PCR), for a number of pathogens [148] and viruses [149–151]. MBs have also demonstrated excellent mismatch discrimination and therefore show application in SNP detection [127,139,144,147,152–154]. In addition, although less common than homogenous solution-based assays, MBs have been immobilized onto DNA microarrays for rapid detection [130,139–141]. 7.7.2 Polymerase Chain Reaction and FRET Used to create thousands–millions of copies of a particular DNA sequence, the PCR has been an indispensable biological tool since its establishment in 1983 [155]. PCR has evolved into a variety of techniques, all based around the basic PCR principle, with a range of applications, including DNA sequencing and gene analysis, forensic studies (such as determining DNA fingerprints), diagnosis of hereditary diseases, and detection of infectious diseases [155]. Over several heating and cooling cycles, DNA is melted [to separate double-stranded deoxyribonucleic acid (dsDNA)] and j289 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.7 Nucleic Acid (DNA/RNA) j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine replicated using each ssDNA as a template, with the help of a DNA polymerase (typically isolated from bacterium Thermus aquaticus called Taq polymerase). Short oligonucleotide sequences known as primers – complementary to the 30 ends of the target DNA – hybridize to the target and initiate amplification in the presence of the DNA polymerase and deoxynucleoside triphosphates. Once the target DNA is replicated, this new strand is used as a template for further replication, allowing the DNA sequence to be amplified by several orders of magnitude. Traditionally, detection was performed after amplification and included colorimetric, chemiluminescent, or gel electrophoresis techniques. However, the advent of closed tube DNA thermal cyclers combined with DNA interchelating dyes (such as ethidium bromide or SYBR Green) or the subsequent FRET-based detection methods for real-time monitoring of PCR amplification have revolutionized the technique by limiting the possibility of contamination, allowing continuous monitoring and accurate quantitation [155]. While the interchelating dyes are simple to use, they lack specificity, detecting both specific and nonspecific amplification. Hence, labeled nucleic acid probes were developed to address this issue [156]. There are now a number of FRET-based probes and primers for quantitative real-time PCR R monitoring, including FRET hybridization probes, TaqMan probes, MB probes, R Snake probes, and Scorpion primers, among others [20,130,133,135,145,157]. In addition, many of these probes–primers, especially the hybridization probes, TaqMan and Scorpion assays, are commercially available, with companies offering a number of common test kits and custom probe design services [see Life Technologies, Integrated DNA Technologies (IDT), and Sigma-Aldrich]. 7.7.2.1 FRET Hybridization Probes In addition to MBs, discussed above, there are two main FRET-based hybridization probe schemes: primer–probe and probe–probe (Figure 7.9b) [134,158]. In the typical case where donor dye–acceptor dye FRET combinations are used, FRET occurs when the PCR cools and the target/probe annealing process begins, resulting in a measured increase in fluorescent signal from the acceptor dye [158]. Subsequently, when the temperature is ramped back up to start the next cycle, the target–probe complex becomes separated and the acceptor dye fluorescence decreases, thus allowing realtime monitoring of PCR. Studying the melting curve of the target/probe and determining the probe melting temperature (Tm) is an excellent method for genotyping [134,158,159]. The probe–probe scheme is generally more popular in melting curve analyses and uses a donor probe, designed to span a specified mutant site on the target DNA sequence, and an acceptor probe, designed to remain annealed to the target DNA sequence (the template) while the donor probe melts. When the donor probe, typically designed for the wild-type target DNA, encounters a mutation in the sequence, its Tm is generally lower than that of the wild-type Tm (for which it is perfectly matched). The extent of Tm variation is dependent upon the type and position of the mismatch, and the neighboring base pairs [134,158]. This assay format forms the basis R technology, and these FRET hybridization probes, of the popular Roche LightCycler combined with melt curve analysis, have been used in a number of clinical applications, including detection of bacteria [160–162], viruses [163,164], fungal species [165], Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 290 and parasites [166], as well as for use in disease diagnostics/prognostics [134,167–170] and forensic analysis [171]. The use of PNA probes has also been proposed as more stable and sensitive alternatives to their DNA analogues [172,173]. 7.7.2.2 TaqMan The TaqMan assay (Figure 7.10) has three components: the template, an energy transfer (ET) probe specific to a certain sequence, and a primer that is needed for PCR amplification [20]. Originally developed by Cetus Corporation [174], the assay is Figure 7.10 The TaqMan assay. The ET probe is complementary to the sequence under investigation and is labeled with a donor– acceptor pair that results in FRET. During the PCR sequence extension stage by Taq polymerase, the ET probe becomes cleaved, separating the donor–acceptor pair and resulting in decreased FRET and increased emission from the donor. Upon completion of the extension reaction, the ET probe is completely dissociated from the target DNA. The donor emission signal increases with increasing PCR cycles. j291 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.7 Nucleic Acid (DNA/RNA) j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine commercially available from Roche (clinical applications) and Life Technologies R (research use), and requires real-time optical instrumentation, such as the COBAS developed by Roche. The ET probe, complementary to the sequence under investigation, has a donor–acceptor pair (in this assay, the donor is referred to as the “Reporter” and the acceptor is typically a quencher species). The primer is located in a region upstream of the probe and is extended by Taq polymerase. As the extension reaction reaches the probe, the probe is cleaved, releasing the donor that emits light. The probe and primer anneal to the template, and only in the presence of a fully complementary sequence will the probe and template completely hybridize. Single point mutations can be detected and are indicated when no fluorescence is detected due to the probe “falling off” the template instead of being cleaved. This assay is favorable because it determines the mutational status in one step during PCR. Despite some studies that claim that traditional SYBR Green and nest PCR are more cost-effective, sensitive, and quicker than the TaqMan assay [175,176], it is routinely used for a number of potential clinical applications such as detection of parasites [176] and a number of viruses [177–179]. 7.7.2.3 Scorpion Assay The Scorpion primer was originally developed by AstraZeneca scientists [180], and is now owned by QIAGEN (via DxS Ltd.). The Scorpion combines the PCR probe and primer into a single molecule, with the probe–primer sections linked by a “blocker” sequence that prevents copying of the probe portion of the Scorpion, which would lead to false-positive signals [20,157]. There are two Scorpion formats known as the stem–loop format and duplex format (also known as, linear format) (Figure 7.11a) [157,181]. In both formats, a dye–quencher FRET pair combination is common with the probe containing the donor fluorophore, which is quenched in the presence of the acceptor quencher attached either on the stem–loop structure (stem–loop format) or on a short complementary hybridized strand (duplex format) [20]. The primer becomes annealed to the target sequence, and is extended via PCR. At the end of the heating and cooling cycles, the temperature is increased, causing the template and extended primer to denature, at which point the probe can hybridize with its target sequence in the extended primer, located downstream of the initial primer sequence, resulting in an increase in the donor fluorescence (Figure 7.11b) for the stem–loop format. There are various advantages and disadvantages to each Scorpion format, although the duplex format is easier to produce and purify, and has a more pronounced FRET response due to larger separation in the distance between quencher and fluorophore upon hybridization to the extended primer [157]. The Scorpion assay is valuable because unimolecular hybridization (primer–probe combination) is faster than bimolecular hybridization (such as TaqMan), and the reagents are commercially available from a number of sources, including QIAGEN, Sigma-Aldrich, and Premier Biosoft International [20,181]. Scorpion primers can distinguish SNPs in a one-step process and are suitable for use with a LightCycler [181,182]. The Scorpion assay has been applied to bacterial, parasitic, and viral detection [182–184], and can be used for point mutation analysis found in many diseases [181,185–187]. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 292 Figure 7.11 The Scorpion assay. (a) The Scorpion combines the PCR probe and primer into a single molecule, with the probe–primer sections linked by a “blocker” sequence that prevents copying of the probe portion of the Scorpion. There are two Scorpion formats shown, the stem–loop format and duplex format (also known as, linear format), both of which typically use a donor/quencher FRET combination. (b) Typical PCR amplification reaction using the stem–loop Scorpion format. Prior to the PCR, the Scorpion probe–primer molecule undergoes efficient FRET. Introduction of the target DNA causes the primer to become annealed and extension is initiated via addition of DNA polymerase and nucleotide bases. At the end of the extension process, the temperature is increased, causing the template and extended primer to denature. This allows the probe portion of the Scorpion molecule to hybridize with its target sequence in the extended primer, located downstream of the initial primer sequence. Subsequently, the FRET is disrupted, resulting in an increase in the donor fluorescence. j293 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.7 Nucleic Acid (DNA/RNA) j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 7.7.2.4 Others There are a couple of other noteworthy examples in the literature of FRET-based probes or primers for use with PCR. Ahmad and Ghasemi sought to create novel FRET primers instead of probes for use in quantitative real-time PCR, taking advantage of the fact that primers are generally specific and are designed to have a low possibility of annealing to other sequences besides the target DNA [135]. In addition, the use of labeled primers avoids the need for additional probes, hence speeding up the PCR. The FRET primers were shown to be more efficient than unlabeled primers detected using the DNA interchelating dye SYBR Green I. PCR has also been monitored in real time using the Snake assay that uses FRETbased probes for detection. The Snake assay combines features of both the TaqMan and Scorpion assays [156]. The assay uses a PCR primer that carries a unique 50 -flap sequence that later causes the PCR amplicon to fold into a stem–loop structure, much like a Scorpion assay. A FRET-based probe then binds the folded PCR amplicon and forms a cleavage structure optimal for 50 nuclease activity, which, like the TaqMan assay, cleaves the FRET probe causing a measurable change in the FRET signal [156]. 7.7.3 Isothermal Amplification Reactions and FRET There are a number of amplification techniques that do not require the thermal cycling necessary for PCR, making them interesting candidates for PoC diagnostics, discussed later. These approaches are referred to as isothermal techniques and include nucleic acid sequence-based amplification (NASBA), helicase-dependent amplification (HDA), recombinase polymerase amplification (RPA), transcriptionmediated amplification (TMA), loop-mediated isothermal amplification (LAMP), rolling circle amplification (RCA), single primer isothermal amplification (SPIA), smart amplification process version 2 (SMAP2), strand displacement amplification (SDA), nicking and extension amplification reaction (NEAR), isothermal chain amplification (ICA), and isothermal and chimeric primer-initiated amplification of nucleic acids (ICAN) [188,189]. A number of these techniques use FRET-based probes for detection. NASBA often combines MB probes with a primer plus a threeenzyme cocktail for RNA detection, and has been demonstrated for SARS-associated coronavirus, human bocavirus, and Aspergillus species detection [149,190–192]. The use of TaqMan probes, in combination with HDA, has been investigated for detection of Vibrio cholerae and B. anthracis [193], while Becton, Dickinson and Company (BD) developed a dual dye-labeled hairpin probe in combination with SDA for Chlamydia trachomatis and Neisseria gonorrhoeae detection [194]. These isothermal techniques typically require a number of specialized primers combined with multiple enzymes, which can increase the assay cost, and some methods suffer from limited specificity and sensitivity. To address the sensitivity issue, Jung et al. proposed a combination of ICA with FRET-based cycling probe technology (CPT) and were able to demonstrate single-copy sensitivity for C. trachomatis [188]. This method, called isothermal target and signaling probe amplification (iTPA) was also used by Kim et al. for detection of Salmonella enterica [195]. The iTPA methodology Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 294 was shown to detect a specific gene found in Salmonella species and was used to distinguish 10 infection-causing Salmonella strains from 40 non-Salmonella strains. R assay is also an isothermal amplification technique that uses a The Invader structure-specific flap endonuclease and two probe oligos, an Invader oligo probe and a FRET-labeled detector probe [130,196–198]. The assay has been made commercially available by Third Wave Technologies (subsidiary of Hologic Inc.) and has been used for SNP genotyping [196,198,199]. The Invader Plus assay [200] combines the Invader assay with PCR and has been used to distinguish between a wild type and a vaccine mutant of the varicella-zoster virus [159]. 7.7.4 Clinical Applications of Nucleic Acid Detection Using FRET As our understanding of the molecular basis of disease continues to advance, there is little doubt that genetic testing technologies will become a fundamental set of tools in the areas of disease diagnosis, oncology, pharmacogenomics (and therefore companion diagnostics) and prognostic applications (determining disease susceptibility) [123,201,202]. FRET-based nucleic acid assays have been used for a number of clinical applications highlighted later. 7.7.4.1 Detection of Pathogens Nucleic acid-based sensing technologies that incorporate FRET for signal transduction have been used to distinguish pathogens from one another, to identify agents used as biological weapons, and to measure helpful organisms that are indicative of good health. The highlighted real-time FRET-based techniques are much quicker than traditional cell culture and can allow more precise characterization of pathogens. Detection of pathogens such as bacteria and viruses, especially genotyping, can aid in diagnosis of disease, which in the case of outbreaks can also link the patient to the source, and help identification of drug-resistant strains, allowing appropriate therapeutic treatment for the patient [136,203]. Rapid detection of biothreat agents is an area of clinical concern and a number of researchers are actively involved in studies to address this need. FRET-based sensing has been employed for genotyping B. anthracis, the bacteria associated with anthrax [193,204]. Yersinia pestis, associated with the bubonic plague, has also been detected using the LightCycler [161]. Life Technologies offers commercial TaqMan kits for measurement of both these bacterial species. Respiratory infections due to bacteria and viruses, in particular influenza viruses, are also common clinical targets of interest, and a number of researchers have used LightCycler (Roche) FRET probes and melting curve analysis to detect types and subtypes of influenza viruses [205,206]. The TaqMan assay has been demonstrated to quantitatively detect avian influenza A virus with a limit of detection of 100 copies/reaction and no false-positive results [178]. Severe acute respiratory syndrome (SARS) coronavirus was measured using MB assays [149], while multiplex RT-PCR (reverse transcriptase–polymerase chain reaction) assays that use FRET probes can be used to detect 13 different respiratory viruses [164]. Mycobacterium j295 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.7 Nucleic Acid (DNA/RNA) j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine tuberculosis (MTB), the bacteria that causes tuberculosis (TB) has been measured with Roche’s commercial diagnostic TaqMan kits as well as via use of the LightCycler combined with hybridization probes [160]. Gastroenteritis can result from viral, bacterial, and parasitic infections, and occurs from exposure to contaminated food, water, or infectious individuals. Life Technologies offer a number of commercial TaqMan kits for detection of common foodborne bacterial pathogens, including the E.coli O157:H7, Campylobacter jejuni, S. enterica, and Listeria monocytogenes. Isothermal amplification techniques have been demonstrated for detection of S. enterica [195] and V. cholerae [193]. Optimized MBs have been used as detection probes, following PCR amplification, for E. coli 0157:H7 [148]. FRET based PCR techniques have been developed for detection of adenoviruses that typically result in infections of the upper respiratory tract, however, serotypes 40 and 41 manifest as gastroenteritis [179]. TaqMan assays have been used for detection and genotyping of genogroup I and II noroviruses [203], while Scorpion probes have been developed for detection of parasites such as Giardia lamblia [184]. Other targets of clinical significance that have been detected using FRET-based nucleic acid assays include human papillomavirus (HPV) [150,183,170,207], hepatitis B virus (HBV) [151,208,209], and HIV. Roche have commercial diagnostic TaqMan kits for detection of HIV, hepatitis C virus (HCV), and HBV, while Hologic offers Invaderbased kits for HPV detection. The Invader and the LightCycler technologies have been used for detection and differentiation of wild-type and vaccine-mutant varicella-zoster viruses (etiologic agent of childhood chicken pox and adult shingles) [163,200]. TaqMan assays have been used to diagnose the reactivation process of viruses such as human cytomegalovirus and human herpesvirus-6, which can occur when an infected person becomes immunocompromised [177]. LightCycler and TaqMan assays have also been used to detect the parasite Toxoplasma gondii [166,176], whereas isothermal amplification techniques have been used for detection of human bocavirus and Aspergillus species detection [190–192]. BD developed a dual dye-labeled hairpin probe in combination with isothermal SDA for C. trachomatis and N. gonorrhoeae detection [194]. Jung et al. proposed a combination of ICA with FRET-based CPT to demonstrate single-copy sensitivity for C. trachomatis [188]. Another very important application of DNA genotyping, from a therapeutic perspective, is the detection and identification of drug-resistant pathogen strains [210]. FRET nucleic acid assays have demonstrated the detection/identification of rifampin-, isoniazid-, and multidrug-resistant MTB [162], clarithromycin-resistant Helicobacter pylori (which causes gastritis) [182], ciprofloxacin-resistant Y. pestis [161], and azole-resistant Candida species (fungal) [165]. Genotyping of HCV has been useful from a clinical perspective, as genotype can influence the clinical outcome, when using current anti-HCV therapies [209]. 7.7.4.2 Prognostic and Diagnostic Applications In addition to pathogen detection, genotyping has been employed for a variety of medical purposes, including prognostic, diagnostic, and therapeutic purposes. FRET probes have been used to identify many polymorphisms associated with illnesses (Table 7.2). However, it is important to note that discovering an Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 296 Table 7.2 Biomarkers detected by FRET genotyping and analyzed using melting curve analysis. Biomarker Disease association Reference Chromosomes 9p21.3 and 4q25 Coronary artery disease and atrial fibrillation Liver metabolism of drugs Myocardial infarction protection Thromboembolism and haemochromatosis Deep vein thrombosis [123] Polymorphisms of CYP2C and CYP2C19 Polymorphism FXIII-A Val34Leu Factor V Leiden Factor V Leiden (G1691A) and prothrombin (G20210A) Varicella-zoster virus vaccine and wild-type strains Newborn screening of b-globin IRGM tetranucleotide promoter oligorepeats HLA-B27 (human leukocyte antigen) TNF-a JAK2 Polymorphisms of endothelin-1, endothelin-2, and endothelin receptor A Polymorphisms of the NOS3 Gene R389G, ADRB1, and CAV1 gene Polymorphisms of interleukin-6 [211] [212] [213–216] [167] Shingles [163] Hemoglobinopathies Immunity Autoimmune diseases Septic shock risk factor Thrombophilia Vascular disease susceptibility [217] [218] [219] [220] [221] [222] Hypertension susceptibility [223–225] Coronary heart disease [226] association between a disease and genetic variation is different from causation. Identifying areas of a genome correlated with a disease simply suggests that this location could have important variants, which could then lead to defining the functional significance of variants. Some cardiac conditions/diseases have strong genetic associations, such as the correlation between coronary artery disease (CAD) and mutations on chromosome 9p21.3 or the strong association between atrial fibrillation and mutations at chromosome 4q25 [123]. The LightCycler has been employed in a number of studies that are looking for genetic associations with cardiac conditions. Arjomand-Nahad et al., for example, looked at polymorphisms of endothelin-1, endothelin-2, and endothelin receptorA,whichmayinfluencesusceptibilitytovasculardiseasessuchashypertension (high blood pressure) and cardiac disease [222]. Jia et al. looked at the relationship between polymorphisms of interleukin-6 (IL-6) and coronary heart disease [226]. Hypertension, a leading cause of cardiac disease, has also been the subject of investigation, with FRET-based assays used to determine if polymorphisms of the NOS3 gene R389 G, ADRB1, or CAV1 gene influence susceptibility to hypertension [223–225]. Shemirani and Muszbek utilized the LightCycler to detect a polymorphism (FXIII-A Val34Leu) associated with protection against myocardial infarction [212]. Blood diseases such as thrombophilia, which increase the risk of thrombosis, and hemoglobinopathies, such as sickle-cell disease, represent another current area of research. Ameziane et al. reported the use of FRET probes to detect mutations j297 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.7 Nucleic Acid (DNA/RNA) j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine (factor V Leiden – G1691A and prothrombin – G20210A) associated with deep vein thrombosis, with a throughput of 72 samples/90 min, significantly faster than the 6 h needed to perform restriction fragment length polymorphism (RFLP) [167]. Detection of factor V Leiden, which may be associated with hereditary thromboembolism and haemochromatosis, has also been demonstrated [213–216]. Thrombophilic mutations of Janus kinase 2 (JAK2) gene [221] and newborn screening of hemoglobinopathies [217] have also been studied using FRET-based detection. Other clinical areas of prognostic interest include genotyping of genes encoding immunity [218], genotyping of TNF-a (tumor necrosis factor-a) (a risk factor for septic shock) [220], and identifying mutations associated with hereditary pancreatitis [227]. FRET-based assays have also been used to investigate human leukocyte antigen (HLA)-B27, which is associated with autoimmune diseases [219] and a1antitrypsin protein deficiency, linked to inherited mutant alleles designated PI Z and PI S, which can lead to obstructive lung disease in adults and liver cirrhosis in children [169]. While the examples above mainly describe the use of genetic biomarkers to predict susceptibility to disease, there are other genomic markers that can be used to make a definitive diagnosis of disease. FRET probes have been used to detect a range of mutations/variations in genomic sequences, which can be used to diagnose diseases [123,130]. The use of genetic differences among individuals has increased the understanding of a number of diseases, such as Crohn’s disease, cancer, asthma, malaria, and heart disease. For example, FRET-based genotyping has been employed to assist in the diagnosis of chronic myeloproliferative disorders (CMPDs), such as chronic myeloid leukemia (CML) [168]. Mutations in the KRAS gene, which encodes the KRAS protein – a GTPase involved in many signal transduction pathways – has been implicated in various types of cancers, including lung and colon cancer, and has been subject to a number of studies using FRET-based signal transduction [152,154]. Clayton et al., for example, discovered that 44% of adenocarcinomas in lungs had one of 7 known mutations in the KRAS oncogene [152]. The quantitative nature of these results, obtained by combining RT-quantitative PCR with ARMS allele-specific amplification, could be used to dictate the analytical sensitivity needed for diagnosis of KRAS mutations. QIAGEN offers a number of Scorpion kits for detection of somatic mutations (acquired rather than germline mutations) found in various oncogenes, including PIK3CA, KRAS, EGFR, and BRAF [185,187]. Somatic point mutations in the GNAS gene have been linked to fibrous dysplasia (FD) of bone/McCune–Albright syndrome (MAS) [172]. 7.7.4.3 Pharmacogenomics and Personalized Medicine The discovery that genomic variation in a patient population can influence drug response (pharmacogenomics) has led to the concept of personalized medicine, that is, tailoring treatment to the individual patient characteristics, which in turn has led to companion diagnostics [3,228]. The use of diagnostics to discriminate between allelic variants for therapeutic purposes is becoming increasingly common and FRET-based detection can be advantageous in this arena. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 298 A prime example of linking genomic analysis to an improved therapeutic output is the dosing of warfarin to treat venous thrombosis [123]. While warfarin is a common treatment for thromboembolism, this drug is associated with adverse effects, has a narrow therapeutic range, and patients demonstrate a wide variation in response. Variability in the effectiveness of the drug can be due to patient age, size, drug use, diet, fitness level, and preexisting conditions. Two gene polymorphisms have been found to be important in the pharmacogenomics of warfarin, specifically the genes encoding the enzymes cytochrome P459 enzyme CYP2C9 and vitamin K epoxide reductase complex subunit 1 VKORC1. In 2007, the United States Food and Drug Administration (FDA) approved updated package labeling for warfarin, which was revised to state that these genetic variations may influence how a patient may respond to the drug [229]. Since then, FDA has increasingly approved certain drugs with companion diagnostic tests, where strong evidence of genetic variation and pharmacogenomics has been demonstrated. For example, Zelboraf (vemurafenib) in combination with the Cobas 4800 BRAF V600 mutation test has been approved for latestage skin melanoma [230]. Op den Buijsch et al. used FRET assays for rapid genotyping of the OATP1B1 polymorphisms A388 G and T521C that have been found to alter the pharmacokinetics of the cholesterol-lowering drug pravastatin and the oncology drug irinotecan, in certain ethnic populations [231,232]. Variability in the CYP2D6 gene, which encodes the cytochrome P450 2D6 enzyme, is found to influence the pharmacokinetics of an analgesic tramadol used to treat moderately severe pain [233]. The uses of a LightCycler and melting curve analysis were shown to be useful in the detection of allelic variants of CYP3A and ABCB1 genes that influence the pharmacokinetics of Tacrolimus, an immunosuppressant prescribed following renal transplantation [234]. 7.8 Aptamers Aptamers are short-chain nucleic acid (RNA and ssDNA) and peptide molecules that bind target analytes with high specificity and selectivity, which are finding application in a number of clinical areas such as recognition of molecules in diagnostic assays [235–239]. They are considered akin to antibodies in terms of their molecular recognition abilities, but also offer some unique advantages, including thermal stability, large-scale production, and low immunogenicity [236,237,239,240]. Some disadvantages of aptamers include binding of nonspecific proteins in complex matrices (such as serum) and metal ion sensitivity [241]. Aptamers can be generated against a wide variety of target analytes, from ions, low molecular weight molecules and proteins through microorganisms and are typically isolated from a combinatorial library by an in vitro procedure called SELEX (systematic evolution of ligands by exponential enrichment) that can be automated [236,242,243]. The potential of aptamers for use in diagnostic and j299 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.8 Aptamers j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine biosensing applications, including FRET-based homogeneous assays, has recently been reviewed [236,237,239]. Due to the large structural rearrangement that often occurs as a result of aptamer target recognition, FRET has proven a particularly useful signal transduction mechanism in aptamer-based assays. Examples of assay formats include sandwich and competitive displacement assays (similar to the antibody equivalents) and molecular aptamer beacons (akin to MBs). The sensitivity of FRETbased aptamer assays is influenced by several factors: affinity for the target, the number of fluorophores incorporated into the aptamer, the proximity of the fluorophore to the quencher, and the purity of the FRET complex after chromatographic separation [244]. One of the most popular targets for aptamer detection is thrombin, a regulator of tumor growth, metastasis, and angiogenesis, in blood serum [245], and a number of FRET-based aptamer assay formats have been demonstrated for this protein. The thrombin aptamer is an example of a guanine-rich ssDNA sequence that forms an intramolecular quadraplex structure upon target binding (reviewed in Ref. [246]). Assay formats for thrombin include the stem–loop aptamer beacon [241], the random coil [247], and a structure-switching signaling assay [248] (Figure 7.12). Recently, some unique FRET materials have been incorporated into FRET-based assays for aptamer-based thrombin detection, including graphene, which acts as a quencher [245,249], quantum dots [249], upconverting phosphors, and carbon nanoparticles [250]. Heyduk and later Lee and coworkers developed an aptamer FRET format somewhat akin to a sandwich antibody in which two aptamers, specific for different epitope regions on the thrombin protein, are labeled with short complementary oligonucleotides modified with the FRET pair molecules [251,252]. Upon aptamer binding to thrombin, the short complementary oligonucleotides become close in proximity, resulting in FRET. Lee and coworkers used a Cy3–Cy5 FRET pair and measured thrombin via singlemolecule photon-burst detection, which has the interesting potential to be combined with microfluidic systems for real-time analysis [252]. Besides thrombin, various FRET-based aptamer assays have been developed for a number of clinically relevant targets, including methylphosphonic acid, a metabolite of several organophosphonic based nerve agents [253]; foot and mouth disease biomarkers [244]; drugs such as cocaine [254,255] and theophylline [256]; adenosine deaminase (ADA, an enzyme needed for purine metabolism) [257]; epithelial marker mucin 1, a biomarker for diagnosis of epithelial cancers [258]; plateletderived growth factor (PDGF), a protein that promotes angiogenesis and regulates cell growth [235,259]; and angiogenin, a protein also linked with angiogenesis [260]. Given the specificity offered by aptamers, multiplexed detection is possible and was recently demonstrated for the detection of cocaine, potassium, and adenosine [261]. Here, the assay takes advantage of the “superquencher” abilities of gold nanoparticles to quench fluorescence from three aptamers labeled with different fluorophores. Binding of the target analyte to the aptamer causes that specific aptamer to become displaced from the Au NP surface, resulting in an increase in fluorescence. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 300 Figure 7.12 Thrombin detection using FRETbased aptamer sensors. (a) Random coil format. DNA-based aptamer undergoes a conformational change upon thrombin binding that brings donor and acceptor (quencher) labels into close proximity, resulting in FRET and reducing donor emission. (b) Aptamer beacon. DNA-based aptamer adopts stem–loop structure that unravels in the presence of thrombin, changing the FRET efficiency and increasing the donor emission. (c) Two-stem duplex assembly. Probe ssDNA sequences, labeled with either donor or acceptor (quencher) species bind complementary portions of the DNA-based aptamer, resulting in FRET. Thrombin binding to the aptamer causes a conformational change that displaces one of the probe sequences, disrupting FRET and increasing the donor emission. (Adapted with permission from Ref. [246].) j301 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.8 Aptamers j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 7.9 High-Throughput and Point-of-Care Devices As demonstrated in the previous sections, FRET has been used to monitor a wide variety of molecular interactions and has been applied to the field of sensing for clinical diagnostics and personalized medicine. Since the vast majority of the FRETbased assays illustrated in this chapter are homogeneous in nature and are often as simple as “mix and detect”, there is great potential for incorporation into highthroughput technologies and/or miniaturized platforms for use in clinical laboratories and ultimately as PoC diagnostic devices [1,146,262]. FRET-based nucleic acid tests have impacted the area of high-throughput technologies for clinical laboratories, as evident from the number of commercial instruments and test kits available that allow real-time detection of target DNA amplification, either by PCR or by isothermal-based techniques. Here, advances in the platform technology, rather than the assay probes themselves, have been the primary driving force behind transitioning to high-throughput methods. Leading examples include the Roche LightCycler 1536 instrument, where the 1536 highdensity multiwall plate enables real-time PCR monitoring of 0.5–2 m l samples, and the BD ViperTM System with XTRTM Technology, based on SDA isothermal amplification, which uses a 96-well plate format and has fully automated sample processing, delivering 736 results/8.5 h. Both technologies use FRET-based nucleic acid assays developed for earlier low- and medium-throughput iterations of the device platforms. Many of the other FRET-based assay formats described in this chapter could readily be combined with standard fluorescent microplate readers (e.g., Tecan, Biotek, and PerkinElmer), which, when combined with liquid handling and robotics platforms, can become high-throughput. While high-throughput technologies suitable for clinical laboratories are highly desirable, the ultimate goal of many IVDs researchers and developers is use in a PoC environment, where real-time results in a clinical setting or at home can reduce the time, sample, and reagent volumes, and hence overall cost of the test. The current status and future of PoC diagnostics was recently described in a comprehensive review by Gubala et al., where the authors highlighted many of the technologies that are enabling IVDs to transition from a clinical laboratory to a PoC environment [1]. The article identified key trends such as personalized medicine, home testing, and multiplexing, and revealed unmet needs for these types of technologies. 7.9.1 PoC Technology Advances FRET-based signal transduction has much to offer, and likewise, has benefited from many of the technologies developed for PoC diagnostic applications such as microfluidics, lab-on-a-chip, and improved optical detection (i.e., better excitation and detection sources) [1,146,262,263]. Microfluidics and lab-on-a-chip technology are key areas for PoC devices, enabling the miniaturization and integration of components (sampling, testing, and detection that comprise the complete assay) to Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 302 create benchtop and portable/handheld instruments [1,264]. Simple “mix-anddetect” FRET assays have been demonstrated for protease detection using a miniaturized fluorimeter, comprising spatial electroluminescent (EL) or light-emitting diode (LED) excitation coupled with CCD (charge-coupled device) detection, for botulinum neurotoxin A or trypsin activity detection [86,98,265]. Results obtained with the miniaturized fluorimeter were comparable to a standard benchtop fluorescence plate reader. There are some commercial handheld fluorimeters (for example, AquaFluorTM and Picofluor by Turner Designs and QuantiFluorTM by Promega) containing UV (ultraviolet), blue, yellow, and/or green LEDs for excitation, but they are typically single sample (cuvette) devices (not high-throughput) and have not yet been applied to FRET-based assays. Nucleic acid detection has been an area of interest for PoCtechnology manufacturers, especially considering potential applications in personalized medicine, where PoC detection could facilitate appropriate and timely treatment [202]. Detection of DNA hybridization has been demonstrated via FRET in microfluidic channels (Figure 7.13) [266–268] and microfluidic droplets, so-called “droplet assays” [269]. In addition, performing traditional PCR in a microfluidic (microPCR) environment is highly desirable [270]. Of all the assay formats described in this chapter, PCR is one of the most complicated to transition to a microfluidic or miniaturized platform, as it involves repeated and precise heating and cooling cycles that can put strain on the materials used and the power requirements of the device [270]. MicroPCR techniques can be classified intotwomaintypes:well-basedandcontinuous-flowchips[270].DigitalPCRrepresents the state-of-the-art in this field and can be either array- or droplet-(a continuous-flow Figure 7.13 FRET-based DNA hybridization detection in microfluidic chips. Schematic illustration of an alligator teeth-shaped PDMS microfluidic chip used for FRET detection of DNA hybridization. Fluorescence emission spectra were recorded at different positions along the microfluidic channel, demonstrating the increase in FRET as increasing amounts of DNA targets and probes hybridize. (Reprinted with permission from Ref. [266]. Copyright, Elsevier.) j303 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 7.9 High-Throughput and Point-of-Care Devices j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine technique) based [186,271,272]. There are a number of commercial benchtop digital PCR instruments available (Fluidigm Corporation, Life Technologies, Bio-Rad Laboratories,andRainDance), whichtypicallyincorporateTaqManand/or Scorpionprobes for detection [186,272,273]. Digital PCR technology has been used to quantify lung cancer EGFR alterations [186] and KRAS mutations [273]. Girkin et al. recently described a miniaturized, fully integrated genotyping system that performed PCR and used FRET-based hybridization probes to carry out melting curve analyses to detect SNPs [274]. This instrument performed an assay in 18 min using both purified DNA and saliva samples contained in capillary tubes with a sample size of 4 ml. Isothermal amplification techniques that do not require the repeated thermal cycles needed for PCR are also emerging for use in combination with microfluidic technologies for PoC applications [189,275]. 7.9.2 PoC Material Advances In addition to advances in technology that incorporate FRET and PoC, substantial progress has been made in the area of new and improved materials for FRET. In particular, nanomaterials can address some of the inherent complications of FRET, that is, photobleaching, spectral cross talk, and direct excitation of the acceptor to propel the wider application of FRET for signal transduction [2]. Of the many types of materials that have the potential for use in FRETapplications (discussed in chapter 6), QDs, lanthanide-based materials, and superquenchers (such as gold and graphene) seem the most common to date for clinical diagnostic applications [276–278]. QDs in particular have found utility in FRET-based assays, primarily as donors, but are increasingly used as acceptors in a number of studies [276,279–283]. QDs possess photophysical properties that make them ideal for FRET applications, particularly those destined for PoC technologies. QDs have narrow photoluminescent (PL) emission profiles that are continuously tunable through control of the QD size and material, and across a broad spectral range, allowing relatively straightforward multiplexing of assays in various formats [147,284]. Broad absorption profiles result in the ability to excite QDs at wavelengths from the UV to the blue side of their emission, resulting in the potential for a large Stokes shift, thus reducing background signals that result from ambient excitation light and direct excitation of the acceptor. The broad absorption profiles suggest that PoC technologies may only require one excitation source for different colored QDs, which is ideal for multiplexing while simplifying PoC instrumentation. QDs typically have high quantum yields and due to their particle nature (i.e., nanoscaffold properties), multiple acceptors can be linked to the QD surface to enhance energy transfer efficiency [279,281]. Such properties can be important for sensor performance, especially when designing PoC devices where there may be trade-offs in terms of excitation power/intensity and device size/weight. Other materials, including lanthanide-based materials and superquenchers, such as Au NPs, carbon nanotubes (CNTs), and graphene sheets, have unique properties that make them useful for energy transfer-based assays. Lanthanide-based materials, which include LLC and lanthanide-doped upconverting phosphors (UCPs), are Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 304 typically used as donors and possess a number of narrow donor emission peaks (Figure 7.5b), with a large Stokes shift [276,277]. They also have long-lived fluorescent lifetimes, making time-resolved, as opposed to steady-state (or frequency), measurements possible. This property of lanthanide-based materials allows timegated measurements that eliminate background fluorescence, thereby improving the signal-to-noise ratio of the sensor [285]. Lanthanide donors have been coupled to a range of acceptor materials, including organic dyes [108,286], carbon nanoparticles R [250], and QDs [29,109,287,288]. TRACE technology, a LLCs-based fluorescent R immunoassay, forms the basis of the detection scheme used by the KRYPTOR series of commercial plate reader systems and was demonstrated for the detection of prostate-specific antigen [108]. So-called “superquenchers” are materials capable of quenching luminescence from a wide range of donor materials, which makes them useful for multiplexing, and they have been demonstrated in a number of FRETbased sensing measurements [245,249,261,289]. 7.10 Conclusions This chapter has demonstrated the huge potential of FRET-based signal transduction for clinical IVD applications and personalized medicine. The intrinsic distance dependence of FRET makes it ideal for monitoring a wide range of molecular recognition events, and detecting many types of target analytes. The diverse array of sensing molecules, ranging from small organic molecules to relatively large polymers, which can be used for FRET, makes this technique very versatile. Nucleic acid, followed by protein-based (e.g., binding proteins, peptide, and antibody) FRET assays, are the main types of FRET-based sensors currently being used, with aptamers being increasingly utilized. Advances in microfluidics, nanotechnology, and materials have the potential to revolutionize FRET-based assay formats by simplifying platform designs. Mainstream use of PoC devices for diagnostic, prognostic, and therapeutic clinical applications, as well as for environmental, defense, and other needs may become a reality in the not too distant future. With further advances in technology, we anticipate FRET to become a more greatly appreciated analytical tool. References 1 Gubala, V., Harris, L.F., Ricco, A.J., Tan, M.X., and Williams, D.E. (2012) Point of care diagnostics: status and future. Analytical Chemistry, 84 (2), 487–515. 2 Coto-Garcia, M.A., Sotelo-Gonzalez, E., Fernandez-Arguelles, M., Pereiro, R., Costa-Fernandez, J.M., and Sanz-Medel, A. (2011) Nanoparticles as fluorescent labels for optical imaging and sensing in genomics and proteomics. Analytical and Bioanalytical Chemistry, 399 (1), 29–42. 3 Sapsford, K.E., Tezak, Z., Kondratovich, M., Pacanowski, M.A., Zineh, I., and Mansfield, E. (2010) Biomarkers to improve the benefit/risk balance for approved therapeutics: a US FDA j305 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 4 5 6 7 8 9 10 11 12 perspective on personalized medicine. Therapeutic Delivery, 1 (5), 631–641. Atkinson, A.J., Colburn, W.A., DeGruttola, V.G., Demets, D.L., Downing, G.J., Hoth, D.F., Oates, J.A., Peck, C.C., Schooley, R.T., Spilker, B.A., Woodcock, J., and Zeger, S.L. (2001) Biomarkers and surrogate endpoints: preferred definitions and conceptual framework. Clinical Pharmacology & Therapeutics, 69 (3), 89–95. Sapsford, K.E., Berti, L., and Medintz, I.L. (2006) Materials for fluorescence resonance energy transfer analysis: beyond traditional donor–acceptor combinations. Angewandte Chemie, International Edition, 45 (28), 4562–4588. Bambot, S.B., Sipior, J., Lakowicz, J.R., and Rao, G. (1994) Lifetime-based optical sensing of pH using resonance energytransfer in sol–gel films. Sensors and Actuators B: Chemical, 22 (3), 181–188. Sipior, J., Bambot, S., Romauld, M., Carter, G.M., Lakowicz, J.R., and Rao, G. (1995) A lifetime-based optical CO2 gas sensor with blue or red excitation and stokes or anti-stokes detection. Analytical Biochemistry, 227 (2), 309–318. Chang, Q., Sipior, J., Lakowicz, J.R., and Rao, G. (1995) A lifetime-based fluorescence resonance energy transfer sensor for ammonia. Analytical Biochemistry, 232 (1), 92–97. Von Bultzingslowen, C., McEvoy, A.K., McDonagh, C., and MacCraith, B.D. (2003) Lifetime-based optical sensor for high-level pCO2 detection employing fluorescence resonance energy transfer. Analytica Chimica Acta, 480 (2), 275–283. Szmacinski, H. and Lakowicz, J.R. (1995) Fluorescence lifetime-based sensing and imaging. Sensors and Actuators B: Chemical, 29 (1–3), 16–24. Ruedas-Rama, M.J. and Hall, E.A. (2010) Analytical nanosphere sensors using quantum dot–enzyme conjugates for urea and creatinine. Analytical Chemistry, 82 (21), 9043–9049. Freeman, R., Li, Y., Tel-Vered, R., Sharon, E., Elbaz, J., and Willner, I. (2009) Selfassembly of supramolecular aptamer structures for optical or electrochemical sensing. Analyst, 134 (4), 653–656. 13 Wang, D., Liu, T., Yin, J., and Liu, S. 14 15 16 17 18 19 20 21 22 (2011) Stimuli-responsive fluorescent poly (N-isopropylacrylamide) microgels labeled with phenylboronic acid moieties as multifunctional ratiometric probes for glucose and temperatures. Macromolecules, 44 (7), 2282–2290. Yin, J., Li, C., Wang, D., and Liu, S. (2010) FRET-derived ratiometric fluorescent Kþ sensors fabricated from thermoresponsive poly(Nisopropylacrylamide) microgels labeled with crown ether moieties. The Journal of Physical Chemistry. B, 114 (38), 12213–12220. Hong, S.W. and Jo, W.H. (2008) A fluorescence resonance energy transfer probe for sensing pH in aqueous solution. Polymer, 49 (19), 4180–4187. Chiu, Y.L., Chen, S.A., Chen, J.H., Chen, K.J., Chen, H.L., and Sung, H.W. (2010) A dual-emission Forster resonance energy transfer nanoprobe for sensing/imaging pH changes in the biological environment. ACS Nano, 4 (12), 7467–7474. Davis, B.W., Niamnont, N., Dillon, R., Bardeen, C.J., Sukwattanasinitt, M., and Cheng, Q. (2011) FRET detection of proteins using fluorescently doped electrospun nanofibers and pattern recognition. Langmuir, 27 (10), 6401–6408. Ambade, A.V., Sandanaraj, B.S., Klaikherd, A., and Thayumanavan, S. (2007) Fluorescent polyelectrolytes as protein sensors. Polymer International, 56 (4), 474–481. Zeng, X., Andrade, C.A., Oliveira, M.D., and Sun, X.L. (2012) Carbohydrate– protein interactions and their biosensing applications. Analytical and Bioanalytical Chemistry, 402 (10), 3161–3176. Sapsford, K.E., Berti, L., and Medintz, I.L. (2004) Fluorescence resonance energy transfer: concepts, applications and advances. Minerva Biotecnologica, 16 (4), 247–273. Song, X.D., Shi, J., Nolan, J., and Swanson, B. (2001) Detection of multivalent interactions through twotiered energy transfer. Analytical Biochemistry, 291 (1), 133–141. Ma, G. and Cheng, Q. (2006) Manipulating FRET with polymeric Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 306 23 24 25 26 27 28 29 30 vesicles: development of a “mix-anddetect” type fluorescence sensor for bacterial toxin. Langmuir, 22 (16), 6743–6745. McGiven, J.A., Thompson, I.J., Commander, N.J., and Stack, J.A. (2009) Time-resolved fluorescent resonance energy transfer assay for simple and rapid detection of anti-Brucella antibodies in ruminant serum samples. Journal of Clinical Microbiology, 47 (10), 3098–3107. Lesch, H.P., Kaikkonen, M.U., Pikkarainen, J.T., and Yla-Herttuala, S. (2010) Avidin–biotin technology in targeted therapy. Expert Opinion on Drug Delivery, 7 (5), 551–564. Wilchek, M., Bayer, E.A., and Livnah, O. (2006) Essentials of biorecognition: the (strept)avidin–biotin system as a model for protein–protein and protein–ligand interaction. Immunology Letters, 103 (1), 27–32. Zempleni, J., Wijeratne, S.S.K., and Hassan, Y.I. (2009) Biotin. Biofactors, 35 (1), 36–46. Liao, F., Xie, Y., Yang, X., Deng, P., Chen, Y., Xie, G., Zhu, S., Liu, B., Yuan, H., Liao, J., Zhao, Y., and Yu, M. (2009) Homogeneous noncompetitive assay of protein via Forster-resonance-energytransfer with tryptophan residue(s) as intrinsic donor(s) and fluorescent ligand as acceptor. Biosensors & Bioelectronics, 25 (1), 112–117. Geissler, D., Charbonniere, L.J., Ziessel, R.F., Butlin, N.G., Loehmannsroeben, H. G., and Hildebrandt, N. (2010) Quantum dot biosensors for ultrasensitive multiplexed diagnostics. Angewandte Chemie, International Edition, 49 (8), 1396–1401. Charbonniere, L.J., Hildebrandt, N., Ziessel, R.F., and Loehmannsroeben, H. G. (2006) Lanthanides to quantum dots resonance energy transfer in timeresolved fluoro-immunoassays and luminescence microscopy. Journal of the American Chemical Society, 128 (39), 12800–12809. Gu, J.Q., Shen, J., Sun, L.D., and Yan, C. H. (2008) Resonance energy transfer in steady-state and time-decay fluoroimmunoassays for lanthanide 31 32 33 34 35 36 37 38 39 nanoparticles based on biotin and avidin affinity. The Journal of Physical Chemistry. C, Nanomaterials and Interfaces, 112 (17), 6589–6593. Oh, E., Hong, M.Y., Lee, D., Nam, S.H., Yoon, H.C., and Kim, H.S. (2005) Inhibition assay of biomolecules based on fluorescence resonance energy transfer (FRET) between quantum dots and gold nanoparticles. Journal of the American Chemical Society, 127 (10), 3270–3271. Ji, E., Wu, D., and Schanze, K.S. (2010) Intercalation-FRET biosensor with a helical conjugated polyelectrolyte. Langmuir, 26 (18), 14427–14429. Saleh, S.M., Mueller, R., Mader, H.S., Duerkop, A., and Wolfbeis, O.S. (2010) Novel multicolor fluorescently labeled silica nanoparticles for interface fluorescence resonance energy transfer to and from labeled avidin. Analytical and Bioanalytical Chemistry, 398 (4), 1615–1623. Liu, L., Wei, G., Liu, Z., He, Z., Xiao, S., and Wang, Q. (2008) Two-photon excitation fluorescence resonance energy transfer with small organic molecule as energy donor for bioassay. Bioconjugate Chemistry, 19 (2), 574–579. Ha, J.H. and Loh, S.N. (2012) Protein conformational switches: from nature to design. Chemistry: A European Journal, 18 (26), 7984–7999. Dwyer, M.A. and Hellinga, H.W. (2004) Periplasmic binding proteins: a versatile superfamily for protein engineering. Current Opinion in Structural Biology, 14 (4), 495–504. Galban, J., Sanz-Vicente, I., Ortega, E., del Barrio, M., and de Marcos, S. (2012) Reagentless fluorescent biosensors based on proteins for continuous monitoring systems. Analytical and Bioanalytical Chemistry, 402 (10), 3039–3054. Pickup, J.C., Hussain, F., Evans, N.D., and Sachedina, N. (2005) In vivo glucose monitoring: the clinical reality and the promise. Biosensors & Bioelectronics, 20 (10), 1897–1902. Hou, B.H., Takanaga, H., Grossmann, G., Chen, L.C., Qu, X.Q., Jones, A.M., Lalonde, S., Schweissgut, O., Wiechert, W., and Frommer, W.B. (2011) Optical j307 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 40 41 42 43 44 45 46 47 48 sensors for monitoring dynamic changes of intracellular metabolite levels in mammalian cells. Nature Protocols, 6 (11), 1818–1833. Fehr, M., Frommer, W.B., and Lalonde, S. (2002) Visualization of maltose uptake in living yeast cells by fluorescent nanosensors. Proceedings of the National Academy of Sciences of the United States of America, 99 (15), 9846–9851. Medintz, I.L. and Deschamps, J.R. (2006) Maltose-binding protein: a versatile platform for prototyping biosensing. Current Opinion in Biotechnology, 17 (1), 17–27. Pickup, J.C., Zhi, Z.L., Khan, F., Saxl, T., and Birch, D.J. (2008) Nanomedicine and its potential in diabetes research and practice. Diabetes Metabolism Research and Reviews, 24 (8), 604–610. Khan, F., Gnudi, L., and Pickup, J.C. (2008) Fluorescence-based sensing of glucose using engineered glucose/ galactose-binding protein: a comparison of fluorescence resonance energy transfer and environmentally sensitive dye labelling strategies. Biochemical and Biophysical Research Communications, 365 (1), 102–106. Garrett, J.R., Wu, X., Jin, S., and Ye, K. (2008) pH-insensitive glucose indicators. Biotechnology Progress, 24 (5), 1085–1089. Veetil, J.V., Jin, S., and Ye, K. (2010) A glucose sensor protein for continuous glucose monitoring. Biosensors & Bioelectronics, 26 (4), 1650–1655. Jin, S., Veetil, J.V., Garrett, J.R., and Ye, K. (2011) Construction of a panel of glucose indicator proteins for continuous glucose monitoring. Biosensors & Bioelectronics, 26 (8), 3427–3431. Dezhurov, S.V., Volkova, I.Y., and Wakstein, M.S. (2011) FRET-based biosensor for oleic acid in nanomolar range with quantum dots as an energy donor. Bioconjugate Chemistry, 22 (3), 338–345. Thurley, S., Roeglin, L., and Seitz, O. (2007) Hairpin peptide beacon: duallabeled PNA-peptide-hybrids for protein detection. Journal of the American Chemical Society, 129 (42), 12693–12695. 49 Medintz, I.L., Goldman, E.R., Lassman, 50 51 52 53 54 55 56 57 M.E., and Mauro, J.M. (2003) A fluorescence resonance energy transfer sensor based on maltose binding protein. Bioconjugate Chemistry, 14 (5), 909–918. Medintz, I.L., Clapp, A.R., Mattoussi, H., Goldman, E.R., Fisher, B., and Mauro, J. M. (2003) Self-assembled nanoscale biosensors based on quantum dot FRET donors. Nature Materials, 2 (9), 630–638. Medintz, I.L., Anderson, G.P., Lassman, M.E., Goldman, E.R., Bettencourt, L.A., and Mauro, J.M. (2004) General strategy for biosensor design and construction employing multifunctional surfacetethered components. Analytical Chemistry, 76 (19), 5620–5629. D’Auria, S., DiCesare, N., Staiano, M., Gryczynski, Z., Rossi, M., and Lakowicz, J.R. (2002) A novel fluorescence competitive assay for glucose determinations by using a thermostable glucokinase from the thermophilic microorganism Bacillus stearothermophilus. Analytical Biochemistry, 303 (2), 138–144. Tolosa, L., Malak, H., Raob, G., and Lakowicz, J.R. (1997) Optical assay for glucose based on the luminescence decay time of the long wavelength dye Cy5TM. Sensors and Actuators B: Chemical, 45 (2), 93–99. Tolosa, L., Szmacinski, H., Rao, G., and Lakowicz, J.R. (1997) Lifetime-based sensing of glucose using energy transfer with a long lifetime donor. Analytical Biochemistry, 250 (1), 102–108. Russell, R.J., Pishko, M.V., Gefrides, C.C., McShane, M.J., and Cote, G.L. (1999) A fluorescence-based glucose biosensor using concanavalin A and dextran encapsulated in a poly(ethylene glycol) hydrogel. Analytical Chemistry, 71 (15), 3126–3132. McCartney, L.J., Pickup, J.C., Rolinski, O. J., and Birch, D.J.S. (2001) Near-infrared fluorescence lifetime assay for serum glucose based on allophycocyanin-labeled concanavalin A. Analytical Biochemistry, 292 (2), 216–221. Liang, F., Pan, T.S., and Sevick-Muraca, E. M. (2005) Measurements of FRET in a glucose-sensitive affinity system with Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 308 58 59 60 61 62 63 64 65 66 67 frequency-domain lifetime spectroscopy. Photochemistry and Photobiology, 81 (6), 1386–1394. Cheung, K.Y., Mak, W.C., and Trau, D. (2008) Reusable optical bioassay platform with permeability-controlled hydrogel pads for selective saccharide detection. Analytica Chimica Acta, 607 (2), 204–210. Peng, J., Wang, Y., Wang, J., Zhou, X., and Liu, Z. (2011) A new biosensor for glucose determination in serum based on upconverting fluorescence resonance energy transfer. Biosensors & Bioelectronics, 28 (1), 414–420. Kim, Y.P., Park, S., Oh, E., Oh, Y.H., and Kim, H.S. (2009) On-chip detection of protein glycosylation based on energy transfer between nanoparticles. Biosensors & Bioelectronics, 24 (5), 1189–1194. Dennis, J.W., Granovsky, M., and Warren, C.E. (1999) Protein glycosylation in development and disease. Bioessays, 21 (5), 412–421. Erker, W., Sdorra, S., and Basche, T. (2005) Detection of single oxygen molecules with fluorescence-labeled hemocyanins. Journal of the American Chemical Society, 127 (42), 14532–14533. Strianese, M., De Martino, F., Pavone, V., Lombardi, A., Canters, G.W., and Pellecchia, C. (2010) A FRET-based biosensor for NO detection. Journal of Inorganic Biochemistry, 104 (6), 619–624. Dennis, A.M., Rhee, W.J., Sotto, D., Dublin, S.N., and Bao, G. (2012) Quantum dot–fluorescent protein FRET probes for sensing intracellular pH. ACS Nano, 6 (4), 2917–2924. Manz, R.A., Hauser, A.E., Hiepe, F., and Radbruch, A. (2005) Maintenance of serum antibody levels. Annual Review of Immunology, 23, 367–386. Sukhanova, A., Susha, A.S., Bek, A., Mayilo, S., Rogach, A.L., Feldmann, J., Oleinikov, V., Reveil, B., Donvito, B., Cohen, J.H., and Nabiev, I. (2007) Nanocrystal-encoded fluorescent microbeads for proteomics: antibody profiling and diagnostics of autoimmune diseases. Nano Letters, 7 (8), 2322–2327. Tian, L. and Heyduk, T. (2009) Antigen peptide-based immunosensors for rapid 68 69 70 71 72 73 74 75 76 detection of antibodies and antigens. Analytical Chemistry, 81 (13), 5218–5225. Lass-Napiorkowska, A., Heyduk, E., Tian, L., and Heyduk, T. (2012) Detection methodology based on target molecule-induced sequence-specific binding to a single-stranded oligonucleotide. Analytical Chemistry, 84 (7), 3382–3389. Golynskiy, M.V., Rurup, W., and Merkx, M. (2010) Antibody detection by using a FRET-based protein conformational switch. ChemBioChem, 11 (16), 2264–2267. Welser, K., Adsley, R., Moore, B.M., Chan, W.C., and Aylott, J.W. (2011) Protease sensing with nanoparticle based platforms. Analyst, 136 (1), 29–41. Carmona, A.K., Juliano, M.A., and Juliano, L. (2009) The use of fluorescence resonance energy transfer (FRET) peptides for measurement of clinically important proteolytic enzymes. Anais da Academia Brasileira de Ciencias, 81 (3), 381–392. Pavan, S. and Berti, F. (2012) Short peptides as biosensor transducers. Analytical and Bioanalytical Chemistry, 402 (10), 3055–3070. Ispas, C.R., Crivat, G., and Andreescu, S. (2012) Review: recent developments in enzyme-based biosensors for biomedical analysis. Analytical Letters, 45 (2–3), 168–186. Banoczi, Z., Alexa, A., Farkas, A., Friedrich, P., and Hudecz, F. (2008) Novel cell-penetrating calpain substrate. Bioconjugate Chemistry, 19 (7), 1375–1381. Prasuhn, D.E., Feltz, A., Blanco-Canosa, J. B., Susumu, K., Stewart, M.H., Mei, B.C., Yakovlev, A.V., Loukov, C., Mallet, J.M., Oheim, M., Dawson, P.E., and Medintz, I. L. (2010) Quantum dot peptide biosensors for monitoring caspase 3 proteolysis and calcium ions. ACS Nano, 4 (9), 5487–5497. Boeneman, K., Mei, B.C., Dennis, A.M., Bao, G., Deschamps, J.R., Mattoussi, H., and Medintz, I.L. (2009) Sensing caspase 3 activity with quantum dot–fluorescent protein assemblies. Journal of the American Chemical Society, 131 (11), 3828–3829. j309 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 77 Karvinen, J., Laitala, V., Makinen, M.L., 78 79 80 81 82 83 84 Mulari, O., Tamminen, J., Hermonen, J., Hurskainen, P., and Hemmila, I. (2004) Fluorescence quenching-based assays for hydrolyzing enzymes. Application of time-resolved fluorometry in assays for caspase, helicase, and phosphatase. Analytical Chemistry, 76 (5), 1429–1436. Medintz, I.L., Clapp, A.R., Brunel, F.M., Tiefenbrunn, T., Uyeda, H., Chang, E.L., Deschamps, J.R., Dawson, P.E., and Mattoussi, H. (2006) Proteolytic activity monitored by fluorescence resonance energy transfer through quantum-dotpeptide conjugates. Nature Materials, 5 (7), 581–589. Sapsford, K.E., Granek, J., Deschamps, J. R., Boeneman, K., Blanco-Canosa, J.B., Dawson, P.E., Susumu, K., Stewart, M.H., and Medintz, I.L. (2011) Monitoring botulinum neurotoxin a activity with peptide-functionalized quantum dot resonance energy transfer sensors. ACS Nano, 5 (4), 2687–2699. Moss, L., Rasmussen, H., Nudelman, R., Dempsey, J., and Williams, J. (2010) Fluorescent substrates useful as highthroughput screening tools for ADAM9. Combinatorial Chemistry & High Throughput Screening, 13 (4), 358–365. Blasi, J., Chapman, E.R., Link, E., Binz, T., Yamasaki, S., Decamilli, P., Sudhof, T.C., Niemann, H., and Jahn, R. (1993) Botulinum neurotoxin A selectively cleaves the synaptic protein SNAP-25. Nature, 365 (6442), 160–163. Schmidt, J.J. and Bostian, K.A. (1997) Endoproteinase activity of type A botulinum neurotoxin: substrate requirements and activation by serum albumin. Journal of Protein Chemistry, 16 (1), 19–26. Grant, S.A., Weilbaecher, C., and Lichlyter, D. (2007) Development of a protease biosensor utilizing silica nanobeads. Sensors and Actuators B: Chemical, 121 (2), 482–489. Shi, L., Rosenzweig, N., and Rosenzweig, Z. (2007) Luminescent quantum dots fluorescence resonance energy transferbased probes for enzymatic activity and enzyme inhibitors. Analytical Chemistry, 79 (1), 208–214. 85 Suzuki, M., Husimi, Y., Komatsu, H., 86 87 88 89 90 91 92 Suzuki, K., and Douglas, K.T. (2008) Quantum dot FRET biosensors that respond to pH, to proteolytic or nucleolytic cleavage, to DNA synthesis, or to a multiplexing combination. Journal of the American Chemical Society, 130 (17), 5720–5725. Sapsford, K.E., Farrell, D., Sun, S., Rasooly, A., Mattoussi, H., and Medintz, I. L. (2009) Monitoring of enzymatic proteolysis on a electroluminescent-CCD microchip platform using quantum dotpeptide substrates. Sensors and Actuators B: Chemical, 139 (1), 13–21. Zauner, T., Berger-Hoffmann, R., Mu^eller, K., Hoffmann, R., and Zuchner, T. (2011) Highly adaptable and sensitive protease assay based on fluorescence resonance energy transfer. Analytical Chemistry, 83 (19), 7356–7363. Oliveira, M., Torquato, R.J., Alves, M.F., Juliano, M.A., Bromme, D., Barros, N.M., and Carmona, A.K. (2010) Improvement of cathepsin S detection using a designed FRET peptide based on putative natural substrates. Peptides, 31 (4), 562–567. Kobayashi, T., Wada, H., Kamikura, Y., Matsumoto, T., Mori, Y., Kaneko, T., Nobori, T., Matsumoto, M., Fujimura, Y., and Shiku, H. (2007) Decreased ADAMTS13 activity in plasma from patients with thrombotic thrombocytopenic purpura. Thrombosis Research, 119 (4), 447–452. Kokame, K., Nobe, Y., Kokubo, Y., Okayama, A., and Miyata, T. (2005) FRETS-VWF73, a first fluorogenic substrate for ADAMTS13 assay. British Journal of Haematology, 129 (1), 93–100. Branchini, B.R., Rosenberg, J.C., Ablamsky, D.M., Taylor, K.P., Southworth, T.L., and Linder, S.J. (2011) Sequential bioluminescence resonance energy transfer-fluorescence resonance energy transfer-based ratiometric protease assays with fusion proteins of firefly luciferase and red fluorescent protein. Analytical Biochemistry, 414 (2), 239–245. Dacres, H., Michie, M., and Trowell, S.C. (2012) Comparison of enhanced bioluminescence energy transfer donors Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 310 93 94 95 96 97 98 99 for protease biosensors. Analytical Biochemistry, 424 (2), 206–210. Gratz, A., Goetz, C., and Jose, J. (2010) A FRET-based microplate assay for human protein kinase CK2, a target in neoplastic disease. Journal of Enzyme Inhibition and Medicinal Chemistry, 25 (2), 234–239. Kaman, W.E., Hulst, A.G., van Alphen, P. T., Roffel, S., van der Schans, M.J., Merkel, T., van Belkum, A., and Bikker, F. J. (2011) Peptide-based fluorescence resonance energy transfer protease substrates for the detection and diagnosis of Bacillus species. Analytical Chemistry, 83 (7), 2511–2517. Cummings, R.T., Salowe, S.P., Cunningham, B.R., Wiltsie, J., Park, Y.W., Sonatore, L.M., Wisniewski, D., Douglas, C.M., Hermes, J.D., and Scolnick, E.M. (2002) A peptide-based fluorescence resonance energy transfer assay for Bacillus anthracis lethal factor protease. Proceedings of the National Academy of Sciences of the United States of America, 99 (10), 6603–6606. Perpetuo, E.A., Juliano, L., Juliano, M.A., Fratelli, F., Prado, S.M., Pimenta, D.C., and Lebrun, I. (2008) Enzymatic profiling of tetanus and botulinum neurotoxins based on vesicle-associated-membrane protein derived fluorogenic substrates. Protein and Peptide Letters, 15 (10), 1100–1106. Pires-Alves, M., Ho, M., Aberle, K.K., Janda, K.D., and Wilson, B.A. (2009) Tandem fluorescent proteins as enhanced FRET-based substrates for botulinum neurotoxin activity. Toxicon, 53 (4), 392–399. Sun, S., Francis, J., Sapsford, K.E., Kostov, Y., and Rasooly, A. (2010) Multiwavelength spatial LED illumination based detector for in vitro detection of botulinum neurotoxin A activity. Sensors and Actuators B: Chemical, 146 (1), 297–306. Kuang, W.F., Chow, L.P., Wu, M.H., and Hwang, L.H. (2005) Mutational and inhibitive analysis of SARS coronavirus 3C-like protease by fluorescence resonance energy transfer-based assays. Biochemical and Biophysical Research Communications, 331 (4), 1554–1559. 100 Ullman, E.F., Schwarzberg, M., and 101 102 103 104 105 106 107 108 Rubenstein, K.E. (1976) Fluorescent excitation transfer immunoassay: a general method for determination of antigens. The Journal of Biological Chemistry, 251 (14), 4172–4178. Wei, Q., Lee, M., Yu, X., Lee, E.K., Seong, G.H., Choo, J., and Cho, Y.W. (2006) Development of an open sandwich fluoroimmunoassay based on fluorescence resonance energy transfer. Analytical Biochemistry, 358 (1), 31–37. Ohiro, Y., Ueda, H., Shibata, N., and Nagamune, T. (2007) Enhanced fluorescence resonance energy transfer immunoassay with improved sensitivity based on the Fab’-based immunoconjugates. Analytical Biochemistry, 360 (2), 266–272. Sasajima, Y., Aburatani, T., Sakamoto, K., and Ueda, H. (2006) Detection of protein tyrosine phosphorylation by open sandwich fluoroimmunoassay. Biotechnology Progress, 22 (4), 968–973. Watanabe, J. and Ishihara, K. (2008) Single step diagnosis system using the FRET phenomenon induced by antibody-immobilized phosphorylcholine group-covered polymer nanoparticles. Sensors and Actuators B: Chemical, 129 (1), 87–93. Kokko, T., Liljenback, T., Peltola, M.T., Kokko, L., and Soukka, T. (2008) Homogeneous dual-parameter assay for prostate-specific antigen based on fluorescence resonance energy transfer. Analytical Chemistry, 80 (24), 9763–9768. Lee, J.S., Joung, H.A., Kim, M.G., and Park, C.B. (2012) Graphene-based chemiluminescence resonance energy transfer for homogeneous immunoassay. ACS Nano, 6 (4), 2978–2983. Wang, R.E., Tian, L., and Chang, Y.H. (2012) A homogeneous fluorescent sensor for human serum albumin. Journal of Pharmaceutical and Biomedical Analysis, 63, 165–169. Kupstat, A., Kumke, M.U., and Hildebrandt, N. (2011) Toward sensitive, quantitative point-of-care testing (POCT) of protein markers: miniaturization of a homogeneous time-resolved fluoroimmunoassay for prostate-specific j311 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 109 110 111 112 113 114 115 116 antigen detection. Analyst, 136 (5), 1029–1035. Chen, M.J., Wu, Y.S., Lin, G.F., Hou, J.Y., Li, M., and Liu, T.C. (2012) Quantum-dotbased homogeneous time-resolved fluoroimmunoassay of alpha-fetoprotein. Analytica Chimica Acta, 741, 100–105. Heyduk, E. and Heyduk, T. (2010) Fluorescent homogeneous immunosensors for detecting pathogenic bacteria. Analytical Biochemistry, 396 (2), 298–303. Heyduk, E., Moxley, M.M., Salvatori, A., Corbett, J.A., and Heyduk, T. (2010) Homogeneous insulin and C-peptide sensors for rapid assessment of insulin and C-peptide secretion by the islets. Diabetes, 59 (10), 2360–2365. Huang, X. and Ren, J. (2011) Gold nanoparticles based chemiluminescent resonance energy transfer for immunoassay of alpha fetoprotein cancer marker. Analytica Chimica Acta, 686 (1), 115–120. Youn, H.J., Terpetschnig, E., Szmacinski, H., and Lakowicz, J.R. (1995) Fluorescence energy transfer immunoassay based on a long-lifetime luminescent metal–ligand complex. Analytical Biochemistry, 232 (1), 24–30. Tan, C., Gajovic-Eichelmann, N., Stocklein, W.F.M., Polzius, R., and Bier, F. F. (2010) Direct detection of 9tetrahydrocannabinol in saliva using a novel homogeneous competitive immunoassay with fluorescence quenching. Analytica Chimica Acta, 658 (2), 187–192. Ohiro, Y., Ueda, H., Shibata, N., and Nagamune, T. (2010) Development of a homogeneous competitive immunoassay for phosphorylated protein antigen based on the enhanced fluorescence resonance energy transfer technology. Journal of Bioscience and Bioengineering, 109 (1), 15–19. Kattke, M.D., Gao, E.J., Sapsford, K.E., Stephenson, L.D., and Kumar, A. (2011) FRET-based quantum dot immunoassay for rapid and sensitive detection of Aspergillus amstelodami. Sensors, 11 (6), 6396–6410. 117 Lichlyter, D.J., Grant, S.A., and Soykan, O. 118 119 120 121 122 123 124 125 126 127 128 (2003) Development of a novel FRET immunosensor technique. Biosensors & Bioelectronics, 19 (3), 219–226. Ko, S. and Grant, S.A. (2003) Development of a novel FRET method for detection of Listeria or Salmonella. Sensors and Actuators B: Chemical, 96 (1–2), 372–378. Ko, S.H. and Grant, S.A. (2006) A novel FRET-based optical fiber biosensor for rapid detection of Salmonella typhimurium. Biosensors & Bioelectronics, 21 (7), 1283–1290. Stringer, R., Schommer, S., Hoehn, D., and Grant, S.A. (2008) Development of an optical biosensor using gold nanoparticles and quantum dots for the detection of porcine reproductive and respiratory syndrome virus. Sensors and Actuators B: Chemical, 134 (2), 427–431. Grant, S.A., Pierce, M.E., Lichlyter, D.J., and Grant, D.A. (2005) Effects of immobilization on a FRET immunosensor for the detection of myocardial infraction. Analytical and Bioanalytical Chemistry, 381 (5), 1012–1018. Metzker, M.L. (2005) Emerging technologies in DNA sequencing. Genome Research, 15 (12), 1767–1776. Knight, J. (2009) Genetics and the general physician: insights, applications and future challenges. Q JM: An International Journal of Medicine, 102 (11), 757–772. Voelkerding, K.V., Dames, S.A., and Durtschi, J.D. (2009) Next-generation sequencing: from basic research to diagnostics. Clinical Chemistry, 55 (4), 641–658. Almal, S.H. and Padh, H. (2012) Implications of gene copy-number variation in health and diseases. Journal of Human Genetics, 57 (1), 6–13. Lupski, J.R. (2007) Structural variation in the human genome. New England Journal of Medicine, 356 (11), 1169–1171. Kim, S. and Misra, A. (2007) SNP genotyping: technologies and biomedical applications. Annual Review of Biomedical Engineering, 9, 289–320. Lewcock, A. (2010) Medicine made to measure. Chemistry World, 7 (7), 56–61. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 312 129 Berti, L., Xie, J., Medintz, I.L., Glazer, A. 130 131 132 133 134 135 136 137 138 139 N., and Mathies, R.A. (2001) Energy transfer cassettes for facile labeling of sequencing and PCR primers. Analytical Biochemistry, 292 (2), 188–197. Didenko, V.V. (2001) DNA probes using fluorescence resonance energy transfer (FRET): designs and applications. Biotechniques, 31 (5), 1106–1116. Epstein, J.R., Biran, I., and Walt, D.R. (2002) Fluorescence-based nucleic acid detection and microarrays. Analytica Chimica Acta, 469 (1), 3–36. Mamotte, C.D.S. (2006) Genotyping of single nucleotide substitutions. The Clinical Biochemist. Reviews/Australian Association of Clinical Biochemists, 27 (1), 63–75. Juskowiak, B. (2011) Nucleic acid-based fluorescent probes and their analytical potential. Analytical and Bioanalytical Chemistry, 399 (9), 3157–3176. Murugesan, G., Kottke-Marchant, K., Ellis, S., Agah, R., and Tubbs, R. (2005) LightTyperTM platform for highthroughput clinical genotyping. Expert Review of Molecular Diagnostics, 5 (3), 457–471. Ahmad, A.I. and Ghasemi, J.B. (2007) New FRET primers for quantitative realtime PCR. Analytical and Bioanalytical Chemistry, 387 (8), 2737–2743. Versalovic, J. and Lupski, J.R. (2002) Molecular detection and genotyping of pathogens: more accurate and rapid answers. Trends in Microbiology, 10 (10), S15–S21. Zhu, J., Lu, Y., Deng, C., Huang, G., Chen, S., Xu, S., Lv, Y., Mitchelson, K., and Cheng, J. (2010) Assessment of fluorescence resonance energy transfer for two-color DNA microarray platforms. Analytical Chemistry, 82 (12), 5304–5312. Matveeva, E.G., Gryczynski, Z., Stewart, D.R., and Gryczynski, I. (2010) Ratiometric FRET-based detection of DNA and micro-RNA on the surface using TIRF detection. Journal of Luminescence, 130 (4), 698–702. Broude, N.E. (2002) Stem–loop oligonucleotides: a robust tool for molecular biology and biotechnology. Trends in Biotechnology, 20 (6), 249–256. 140 Kim, H., Kane, M.D., Kim, S., 141 142 143 144 145 146 147 148 149 Dominguez, W., Applegate, B.M., and Savikhin, S. (2007) A molecular beacon DNA microarray system for rapid detection of E. coli O157:H7 that eliminates the risk of a false negative signal. Biosensors & Bioelectronics, 22 (6), 1041–1047. Lee, K., Rouillard, J.M., Kim, B.G., Gulari, E., and Kim, J. (2009) Conjugated polymers combined with a molecular beacon for label-free and self-signalamplifying DNA microarrays. Advanced Functional Materials, 19 (20), 3317–3325. Tyagi, S. and Kramer, F.R. (1996) Molecular beacons: probes that fluoresce upon hybridization. Nature Biotechnology, 14 (3), 303–308. Tyagi, S., Marras, S.A.E., and Kramer, F.R. (2000) Wavelength-shifting molecular beacons. Nature Biotechnology, 18 (11), 1191–1196. Li, Y., Zhou, X., and Ye, D. (2008) Molecular beacons: an optimal multifunctional biological probe. Biochemical and Biophysical Research Communications, 373 (4), 457–461. Guo, J., Ju, J., and Turro, N.J. (2012) Fluorescent hybridization probes for nucleic acid detection. Analytical and Bioanalytical Chemistry, 402 (10), 3115–3125. Varghese, S.S., Zhu, Y., Davis, T.J., and Trowell, S.C. (2010) FRET for lab-on-achip devices: current trends and future prospects. Lab on a Chip, 10 (11), 1355–1364. Kim, J.H., Chaudhary, S., and Ozkan, M. (2007) Multicolour hybrid nanoprobes of molecular beacon conjugated quantum dots: FRET and gel electrophoresis assisted target DNA detection. Nanotechnology, 18 (19), 195105. Ram, S., Singh, R.L., and Shanker, R. (2008) In silico comparison of real-time PCR probes for detection of pathogens. In Silico Biology, 8 (3–4), 251–259. Keightley, M.C., Sillekens, P., Schippers, W., Rinaldo, C., and St George, K. (2005) Real-time NASBA detection of SARSassociated coronavirus and comparison with real-time reverse transcription-PCR. j313 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 150 151 152 153 154 155 156 Journal of Medical Virology, 77 (4), 602–608. Takacs, T., Jeney, C., Kovacs, L., Mozes, J., Benczik, M., and Sebe, A. (2008) Molecular beacon-based real-time PCR method for detection of 15 high-risk and 5 low-risk HPV types. Journal of Virological Methods, 149 (1), 153–162. Yates, S., Penning, M., Goudsmit, J., Frantzen, I., van de Weijer, B., van Strijp, D., and van Gemen, B. (2001) Quantitative detection of hepatitis B virus DNA by real-time nucleic acid sequence-based amplification with molecular beacon detection. Journal of Clinical Microbiology, 39 (10), 3656–3665. Clayton, S.J., Scott, F.M., Walker, J., Callaghan, K., Haque, K., Liloglou, T., Xinarianos, G., Shawcross, S., Ceuppens, P., Field, J.K., and Fox, J.C. (2000) K-ras point mutation detection in lung cancer: comparison of two approaches to somatic mutation detection using ARMS allelespecific amplification. Clinical Chemistry, 46 (12), 1929–1938. Lin, Y.W., Ho, H.T., Huang, C.C., and Chang, H.T. (2008) Fluorescence detection of single nucleotide polymorphisms using a universal molecular beacon. Nucleic Acids Research, 36 (19), e123. Wabuyele, M.B., Farquar, H., Stryjewski, W., Hammer, R.P., Soper, S.A., Cheng, Y. W., and Barany, F. (2003) Approaching real-time molecular diagnostics: singlepair fluorescence resonance energy transfer (spFRET) detection for the analysis of low abundant point mutations in K-ras oncogenes. Journal of the American Chemical Society, 125 (23), 6937–6945. Bartlett, J.M.S. and Stirling, D. (2003) A short history of the polymerase chain reaction. Methods in Molecular Biology (Clifton, N. J.), 226, 3–6. Kutyavin, I.V. (2010) New approach to real-time nucleic acids detection: folding polymerase chain reaction amplicons into a secondary structure to improve cleavage of Forster resonance energy transfer probes in 50 -nuclease assays. Nucleic Acids Research, 38 (5), e29. 157 Carters, R., Ferguson, J., Gaut, R., 158 159 160 161 162 163 164 165 Ravetto, P., Thelwell, N., and Whitcombe, D. (2008) Design and use of Scorpions fluorescent signaling molecules. Methods in Molecular Biology (Clifton, N. J.), 429, 99–115. Bernard, P.S. and Wittwer, C.T. (2000) Homogeneous amplification and variant detection by fluorescent hybridization probes. Clinical Chemistry, 46 (2), 147–148. Breen, G. (2002) Novel and alternate SNP and genetic technologies. Psychiatric Genetics, 12 (2), 83–88. Torres, M.J., Criado, A., Palomares, J.C., and Aznar, J. (2000) Use of real-time PCR and fluorimetry for rapid detection of rifampin and isoniazid resistanceassociated mutations in Mycobacterium tuberculosis. Journal of Clinical Microbiology, 38 (9), 3194–3199. Lindler, L.E., Fan, W., and Jahan, N. (2001) Detection of ciprofloxacin-resistant Yersinia pestis by fluorogenic PCR using the LightCycler. Journal of Clinical Microbiology, 39 (10), 3649–3655. Ali, A., Hasan, Z., Moatter, T., Tanveer, M., and Hasan, R. (2009) M. tuberculosis Central Asian Strain 1 MDR isolates have more mutations in rpoB and katG genes compared with other genotypes. Scandinavian Journal of Infectious Diseases, 41 (1), 37–44. Loparev, V.N., McCaustland, K., Holloway, B.P., Krause, P.R., Takayama, M., and Schmid, D.S. (2000) Rapid genotyping of varicella-zoster virus vaccine and wildtype strains with fluorophore-labeled hybridization probes. Journal of Clinical Microbiology, 38 (12), 4315–4319. Lassauniere, R., Kresfelder, T., and Venter, M. (2010) A novel multiplex real-time RTPCR assay with FRET hybridization probes for the detection and quantitation of 13 respiratory viruses. Journal of Virological Methods, 165 (2), 254–260. Loeffler, J., Hagmeyer, L., Hebart, H., Henke, N., Schumacher, U., and Einsele, H. (2000) Rapid detection of point mutations by fluorescence resonance energy transfer and probe melting curves in Candida species. Clinical Chemistry, 46 (5), 631–635. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 314 166 Yang, W., Lindquist, H., Cama, V., 167 168 169 170 171 172 173 Schaefer, F.W., Villegas, E., Fayer, R., Lewis, E.J., Feng, Y., and Xiao, L. (2009) Detection of Toxoplasma gondii oocysts in water sample concentrates by real-time PCR. Applied and Environmental Microbiology, 75 (11), 3477–3483. Ameziane, N., Lamotte, M., Lamoril, J., Lebret, D., Deybach, J.C., Kaiser, T., and de Prost, D. (2003) Combined factor V Leiden (G1691A) and prothrombin (G20210A) genotyping by multiplex realtime polymerase chain reaction using fluorescent resonance energy transfer hybridization probes on the Rotor-Gene 2000. Blood Coagulation & Fibrinolysis, 14 (4), 421–424. Murugesan, G., Aboudola, S., Szpurka, H., Verbic, M.A., Maciejewski, J.P., Tubbs, R.R., and Hsi, E.D. (2006) Identification of the JAK2 V617F mutation in chronic myeloproliferative disorders using FRET probes and melting curve analysis. American Journal of Clinical Pathology, 125 (4), 625–633. Von Ahsen, N., Oellerich, M., and Schutz, E. (2000) Use of two reporter dyes without interference in a single-tube rapid-cycle PCR: a1-antitrypsin genotyping by multiplex real-time fluorescence PCR with the LightCycler. Clinical Chemistry, 46 (2), 156–161. Weissenborn, S.J., Wieland, U., Junk, M., and Pfister, H. (2010) Quantification of beta-human papillomavirus DNA by realtime PCR. Nature Protocols, 5 (1), 1–13. Halpern, M.D. and Ballantyne, J. (2009) A single nucleotide polymorphism melt curve assay employing an intercalating dye probe fluorescence resonance energy transfer for forensic analysis. Analytical Biochemistry, 391 (1), 1–10. Karadag, A., Riminucci, M., Bianco, P., Cherman, N., Kuznetsov, S.A., Nguyen, N., Collins, M.T., Robey, P.G., and Fisher, L.W. (2004) A novel technique based on a PNA hybridization probe and FRET principle for quantification of mutant genotype in fibrous dysplasia/McCune– Albright syndrome. Nucleic Acids Research, 32 (7), e63. Oquare, B.Y. and Taylor, J.S. (2008) Synthesis of peptide nucleic acid FRET 174 175 176 177 178 179 probes via an orthogonally protected building block for post-synthetic labeling of peptide nucleic acids at the 5-position of uracil. Bioconjugate Chemistry, 19 (11), 2196–2204. Holland, P.M., Abramson, R.D., Watson, R., and Gelfand, D.H. (1991) Detection of specific polymerase chain-reaction product by utilizing the 50 -30 exonuclease activity of Thermus aquaticus DNA polymerase. Proceedings of the National Academy of Sciences of the United States of America, 88 (16), 7276–7280. Calero, O., Hortigueela, R., Bullido, M. J., and Calero, M. (2009) Apolipoprotein E genotyping method by real time PCR, a fast and cost-effective alternative to R and FRET assays. Journal the TaqMan of Neuroscience Methods, 183 (2), 238–240. Calderaro, A., Piccolo, G., Gorrini, C., Peruzzi, S., Zerbini, L., Bommezzadri, S., Dettori, G., and Chezzi, C. (2006) Comparison between two real-time PCR assays and a nested-PCR for the detection of Toxoplasma gondii. Acta Bio-medica: Atenei Parmensis, 77 (2), 75–80. Pradeau, K., Couty, L., Szelag, J.C., Turlure, P., Rolle, F., Ferrat, P., Bordessoule, D., Le Meur, Y., Denis, F., and Ranger-Rogez, S. (2006) Multiplex real-time PCR assay for simultaneous quantitation of human cytomegalovirus and herpesvirus-6 in polymorphonuclear and mononuclear cells of transplant recipients. Journal of Virological Methods, 132 (1–2), 77–84. Chantratita, W., Sukasem, C., Kaewpongsri, S., Srichunrusami, C., Pairoj, W., Thitithanyanont, A., Chaichoune, K., Ratanakron, P., Songserm, T., Damrongwatanapokin, S., and Landt, O. (2008) Qualitative detection of avian influenza A (H5N1) viruses: a comparative evaluation of four real-time nucleic acid amplification methods. Molecular and Cellular Probes, 22 (5–6), 287–293. Jothikumar, N., Cromeans, T.L., Hill, V.R., Lu, X.Y., Sobsey, M.D., and Erdman, D.D. (2005) Quantitative real-time PCR assays for detection of human adenoviruses and identification of serotypes 40 and 41. j315 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 180 181 182 183 184 185 186 187 Applied and Environmental Microbiology, 71 (6), 3131–3136. Whitcombe, D., Theaker, J., Guy, S.P., Brown, T., and Little, S. (1999) Detection of PCR products using self-probing amplicons and fluorescence. Nature Biotechnology, 17 (8), 804–807. Solinas, A., Brown, L.J., McKeen, C., Mellor, J.M., Nicol, J.T.G., Thelwell, N., and Brown, T. (2001) Duplex Scorpion primers in SNP analysis and FRET applications. Nucleic Acids Research, 29 (20), e96. Burucoa, C., Garnier, M., Silvain, C., and Fauchere, J.L. (2008) Quadruplex realtime PCR assay using allele-specific Scorpion primers for detection of mutations conferring clarithromycin resistance to Helicobacter pylori. Journal of Clinical Microbiology, 46 (7), 2320–2326. Hart, K.W., Williams, O.M., Thelwell, N., Flander, A.N., Brown, T., Borysiewicz, L. K., and Gelder, C.M. (2001) Novel method for detection, typing, and quantification of human papillomaviruses in clinical samples. Journal of Clinical Microbiology, 39 (9), 3204–3212. Ng, C.T., Gilchrist, C.A., Lane, A., Roy, S., Haque, R., and Houpt, E.R. (2005) Multiplex real-time PCR assay using Scorpion probes and DNA capture for genotype-specific detection of Giardia lamblia on fecal samples. Journal of Clinical Microbiology, 43 (3), 1256–1260. Board, R.E., Thelwell, N.J., Ravetto, P.F., Little, S., Ranson, M., Dive, C.L., Hughes, A., and Whitcombe, D. (2008) Multiplexed assays for detection of mutations in PIK3CA. Clinical Chemistry, 54 (4), 757–760. Wang, J., Ramakrishnan, R., Tang, Z., Fan, W., Kluge, A., Dowlati, A., Jones, R. C., and Ma, P.C. (2010) Quantifying EGFR alterations in the lung cancer genome with nanofluidic digital PCR arrays. Clinical Chemistry, 56 (4), 623–632. Nordgard, O., Oltedal, S., Janssen, E.A., Gilje, B., Korner, H., Tjensvoll, K., and Smaaland, R. (2012) Comparison of a PNA clamp PCR and an ARMS/Scorpion PCR assay for the detection of K-ras mutations. Diagnostic Molecular Pathology, 21 (1), 9–13. 188 Jung, C., Chung, J.W., Kim, U.O., Kim, 189 190 191 192 193 194 195 M.H., and Park, H.G. (2010) Isothermal target and signaling probe amplification method, based on a combination of an isothermal chain amplification technique and a fluorescence resonance energy transfer cycling probe technology. Analytical Chemistry, 82 (14), 5937–5943. Craw, P. and Balachandran, W. (2012) Isothermal nucleic acid amplification technologies for point-of-care diagnostics: a critical review. Lab on a Chip, 12 (14), 2469–2486. Compton, J. (1991) Nucleic acid sequencebased amplification. Nature, 350 (6313), 91–92. Boehmer, A., Schildgen, V., Luesebrink, J., Ziegler, S., Tillmann, R.L., Kleines, M., and Schildgen, O. (2009) Novel application for isothermal nucleic acid sequence-based amplification (NASBA). Journal of Virological Methods, 158 (1–2), 199–201. Park, C., Kwon, E.Y., Shin, N.Y., Choi, S. M., Kim, S.H., Park, S.H., Lee, D.G., Choi, J.H., and Yoo, J.H. (2011) Evaluation of nucleic acid sequence based amplification using fluorescence resonance energy transfer (FRET-NASBA) in quantitative detection of Aspergillus 18S rRNA. Medical Mycology, 49 (1), 73–79. Tong, Y., Tang, W., Kim, H.J., Pan, X., Ranalli, T.A., and Kong, H. (2008) Development of isothermal TaqMan assays for detection of biothreat organisms. Biotechniques, 45 (5), 543–557. Little, M.C., Andrews, J., Moore, R., Bustos, S., Jones, L., Embres, C., Durmowicz, G., Harris, J., Berger, D., Yanson, K., Rostkowski, C., Yursis, D., Price, J., Fort, T., Walters, A., Collis, M., Llorin, O., Wood, J., Failing, F., O’Keefe, C., Scrivens, B., Pope, B., Hansen, T., Marino, K., Williams, K., and Boenisch, M. (1999) Strand displacement amplification and homogeneous real-time detection incorporated in a secondgeneration DNA probe system, BDProbeTecET. Clinical Chemistry, 45 (6), 777–784. Kim, J., Jahng, M., Lee, G., Lee, K., Chae, H., Lee, J., Lee, J., and Kim, M. (2011) Rapid and simple detection of the invA Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 316 196 197 198 199 200 201 202 203 gene in Salmonella spp. by isothermal target and probe amplification (iTPA). Letters in Applied Microbiology, 52 (4), 399–405. R Olivier, M. (2005) The Invader assay for SNP genotyping. Mutation Research, 573 (1–2), 103–110. Mein, C.A., Barratt, B.J., Dunn, M.G., Siegmund, T., Smith, A.N., Esposito, L., Nutland, S., Stevens, H.E., Wilson, A.J., Phillips, M.S., Jarvis, N., Law, S., de Arruda, M., and Todd, J.A. (2000) Evaluation of single nucleotide polymorphism typing with Invader on PCR amplicons and its automation. Genome Research, 10 (3), 330–343. Fors, L., Lieder, K.W., Vavra, S.H., and Kwiatkowski, R.W. (2000) Large-scale SNP scoring from unamplified genomic DNA. Pharmacogenomics, 1 (2), 219–229. Ledford, M., Friedman, K.D., Hessner, M. J., Moehlenkamp, C., Williams, T.M., and Larson, R.S. (2000) A multi-site study for detection of the factor V (Leiden) mutation from genomic DNA using a homogeneous Invader microtiter plate fluorescence resonance energy transfer (FRET) assay. Journal of Molecular Diagnostics, 2 (2), 97–104. Tang, Y.W., Allawi, H.T., DeLeon-Carnes, M., Li, H., Day, S.P., and Schmid, D. (2007) Detection and differentiation of wild-type and vaccine mutant varicellaR zoster viruses using an Invader Plus method. Journal of Clinical Virology, 40 (2), 129–134. Eibl, F. and Prieur, R. (2011) Tools for molecular diagnostics. IVD Technology. http://www.ivdtechnology.com/article/ tools-molecular-diagnostics-11 (accessed 31 August 2011). Ziegler, A., Koch, A., Krockenberger, K., and Grosshennig, A. (2012) Personalized medicine using DNA biomarkers: a review. Human Genetics, 131 (10), 1627–1638. Schultz, A.C., Vega, E., Dalsgaard, A., Christensen, L.S., Norrung, B., Hoorfar, J., and Vinje, J. (2011) Development and evaluation of novel one-step TaqMan realtime RT-PCR assays for the detection and direct genotyping of genogroup I and 204 205 206 207 208 209 210 211 212 II noroviruses. Journal of Clinical Virology, 50 (3), 230–234. Patra, G., Williams, L.E., Qi, Y., Rose, S., Redkar, R., and DelVecchio, V.G. (2002) Rapid genotyping of Bacillus anthracis strains by real-time polymerase chain reaction. Annals of the New York Academy of Sciences, 969, 106–111. Poddar, S.K. (2003) Detection of type and subtypes of influenza virus by hybrid formation of FRET probe with amplified target DNA and melting temperature analysis. Journal of Virological Methods, 108 (2), 157–163. Smith, A.B., Mock, V., Melear, R., Colarusso, P., and Willis, D.E. (2003) Rapid detection of influenza A and B viruses in clinical specimens by Light Cycler real time RT-PCR. Journal of Clinical Virology, 28 (1), 51–58. Kocjan, B.J., Seme, K., and Poljak, M. (2008) Detection and differentiation of human papillomavirus genotypes HPV-6 and HPV-11 by FRET-based real-time PCR. Journal of Virological Methods, 153 (2), 245–249. Wang, X., Lou, X., Wang, Y., Guo, Q., Fang, Z., Zhong, X., Mao, H., Jin, Q., Wu, L., Zhao, H., and Zhao, J. (2010) QDsDNA nanosensor for the detection of hepatitis B virus DNA and the single-base mutants. Biosensors & Bioelectronics, 25 (8), 1934–1940. Germer, J.J., Majewski, D.W., Yung, B., Mitchell, P.S., and Yao, J.D.C. (2006) Evaluation of the Invader assay for genotyping hepatitis C virus. Journal of Clinical Microbiology, 44 (2), 318–323. Garcia-Alvarez, L., Dawson, S., Cookson, B., and Hawkey, P. (2012) Working across the veterinary and human health sectors. Journal of Antimicrobial Chemotherapy, 67 (1), 137–149. Pachkoria, K., Lucena, M., Ruiz-Cabello, I., Crespo, F., Cabello, E., and Andrade, M.R. (2007) Genetic polymorphisms of CYP2C9 and CYP2C19 are not related to drug-induced idiosyncratic liver injury (DILI). British Journal of Pharmacology, 150 (6), 808–815. Shemirani, A.H. and Muszbek, L. (2004) Rapid detection of the factor XIII Val34Leu (163 G!T) polymorphism by j317 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 213 214 215 216 217 218 219 real-time PCR using fluorescence resonance energy transfer detection and melting curve analysis. Clinical Chemistry and Laboratory Medicine, 42 (8), 877–879. Neoh, S.H., Brisco, M.J., Firgaira, F.A., Trainor, K.J., Turner, D.R., and Morley, A. A. (1999) Rapid detection of the factor V Leiden (1691 G > A) and haemochromatosis (845 G > A) mutation by fluorescence resonance energy transfer (FRET) and real time PCR. Journal of Clinical Pathology, 52 (10), 766–769. Martinez-Serra, J., Gutierrez, A., Amat, J., Galmes, B., Vila, A., Julia, M., Lopez, H., Bautista, A., Truyols, C., Ros, T., Navarro, M., Canaro, M., Diaz, M., and Besalduch, J. (2009) Rapid triplex asymmetric real-time PCR hybridization probe assay for the joint genotyping of F2, F5 and F12. Clinical Biochemistry, 42 (12), 1317–1324. Moyses, C., Moreira, E., Asprino, P., Guimaraes, G., and Alberto, F. (2008) Simultaneous detection of the C282Y, H63D and S65C mutations in the hemochromatosis gene using quenchedFRET real-time PCR. Brazilian Journal of Medical and Biological Research, 41 (10), 833–838. Parks, S.B., Popovich, B.W., and Press, R. D. (2001) Real-time polymerase chain reaction with fluorescent hybridization probes for the detection of prevalent mutations causing common thrombophilic and iron overload phenotypes. American Journal of Clinical Pathology, 115 (3), 439–447. Lin, Z.L., Suzow, J.G., Fontaine, J.M., and Naylor, E.W. (2004) A high throughput beta-globin genotyping method by multiplexed melting temperature analysis. Molecular Genetics and Metabolism, 81 (3), 237–243. Intemann, C.D., Thye, T., Sievertsen, J., Owusu-Dabo, E., Horstmann, R.D., and Meyer, C.G. (2009) Genotyping of IRGM tetranucleotide promoter oligorepeats by fluorescence resonance energy transfer. Biotechniques, 46 (1), 58–60. Faner, R., Casamitjana, N., Colobran, R., Ribera, A., Pujol-Borrell, R., Palou, E., and Juan, M. (2004) HLA-B27 genotyping by fluorescent resonance emission transfer 220 221 222 223 224 225 226 (FRET) probes in real-time PCR. Human Immunology, 65 (8), 826–838. Hayashi, N., Imamura, Y., Hiyoshi, Y., Takamori, H., Beppu, T., Hirota, M., and Baba, H. (2008) Rapid genotyping of tumor necrosis factor alpha with fluorogenic hybridization probes on the LightCycler. Clinical and Experimental Medicine, 8 (4), 217–224. Martinez-Serra, J., Maffiotte, E., Gutierrez, A., Duran, M., Amat, J., and Besalduch, J. (2009) New real-time PCRbased method for the joint genotyping of JAK2 (V617F) with inherited thrombophilic F5 and F2 mutations. Clinica Chimica Acta, 410 (1–2), 59–63. Arjomand-Nahad, F., Landt, O., Stangl, K., Diefenbach, K., and Roots, I. (2006) Fifteen polymorphisms in endothelin-1, endothelin-2 and endothelin-receptor-A genotyped by four duplex assays and seven simple assays on a LightCycler using hybridization probes. Clinical Chemistry and Laboratory Medicine, 44 (8), 929–932. Fernandez, M.L., Ruiz, R., Gonzalez, M. A., Ramirez-Lorca, R., Couto, C., Ramos, A., Gutierrez-Tous, R., Rivera, J.M., Ruiz, A., Real, L.M., and Grilo, A. (2004) Association of NOS3 gene with metabolic syndrome in hypertensive patients. Thrombosis and Haemostasis, 92 (2), 413–418. Sauer, M., Tews, J., Schwarzer, M., Seeburger, J., Adams, V., Falk, V., Dhein, S., and Mohr, F.W. (2009) Rapid detection of the R389 G polymorphism of ADRB1 by real time PCR using fluorescence resonance energy transfer. Clinical Laboratory, 55 (3–4), 128–136. Grilo, A., Fernandez, M.L., Beltran, M., Ramirez-Lorca, R., Gonzalez, M.A., Royo, J.L., Gutierrez-Tous, R., Moron, F.J., Couto, C., Serrano-Rios, M., Saez, M.E., Ruiz, A., and Real, L.M. (2006) Genetic analysis of CAVI gene in hypertension and metabolic syndrome. Thrombosis and Haemostasis, 95 (4), 696–701. Jia, X., Tian, Y., Wang, Y., Deng, X., Dong, Z., Scafa, N., and Zhang, X. (2010) Association between the interleukin-6 gene-572G/C and 597G/A polymorphisms and coronary heart Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 318 227 228 229 230 231 232 233 disease in the Han Chinese. Medical Science Monitor, 16 (3), CR103–CR108. Felderbauer, P., Schnekenburger, J., Lebert, R., Bulut, K., Parry, M., Meister, T., Schick, V., Schmitz, F., Domschke, W., and Schmidt, W. (2008) A novel A121T mutation in human cationic trypsinogen associated with hereditary pancreatitis: functional data indicating a loss-offunction mutation influencing the R122 trypsin cleavage site. Journal of Medical Genetics, 45 (8), 507–512. Koch, W.H. (2004) Technology platforms for pharmacogenomic diagnostic assays. Nature Reviews Drug Discovery, 3 (9), 749–761. FDA (2007) FDA approves updated warfarin (Coumadin) prescribing information. New genetic information may help providers improve initial dosing estimates of the anticoagulant for individual patients. Press release, August 16, http://www.fda.gov/NewsEvents/ Newsroom/PressAnnouncements/2007/ ucm108967.htm. FDA (2011) FDA approves Zelboraf and companion diagnostic test for late-stage skin cancer. Press release, August 17, http://www.fda.gov/NewsEvents/ Newsroom/PressAnnouncements/ ucm268241.htm. Op den Buijsch, R.A.M., Wijnen, P.A.H. M., van Dieijen-Visser, M.P., de Vries, J. E., and Bekers, O. (2005) Rapid genotyping of the OATP1B1 polymorphisms A388 G and T521C with real-time PCR FRET assays. Pharmacogenomics, 6 (4), 393–397. Aklillu, E., Mugusi, S., Ngaimisi, E., Hoffmann, M.M., Koenig, S., Ziesenitz, V., Mikus, G., Haefeli, W.E., and Weiss, J. (2011) Frequency of the SLCO1B1 388A>G and the 521T>C polymorphism in Tanzania genotyped by a new R -based method. European LightCycler Journal of Clinical Pharmacology, 67 (11), 1139–1145. Borlak, J., Hermann, R., Erb, K., and Thum, T. (2003) A rapid and simple CYP2D6 genotyping assay: case study with the analgetic tramadol. Metabolism: Clinical and Experimental, 52 (11), 1439–1443. 234 Cheung, C.Y., den Buijsch, R.A.M.O., 235 236 237 238 239 240 241 242 243 244 Wong, K.M., Chan, H.W., Chau, K.F., Li, C.S., Leung, K.T., Kwan, T.H., de Vrie, J. E., Wijnen, P.A.H.M., van Dieen-Visser, M.P., and Bekers, O. (2006) Influence of different allelic variants of the CYP3A and ABCB1 genes on the tacrolimus pharmacokinetic profile of Chinese renal transplant recipients. Pharmacogenomics, 7 (4), 563–574. Kim, G.I., Kim, K.W., Oh, M.K., and Sung, Y.M. (2009) The detection of platelet derived growth factor using decoupling of quencher-oligonucleotide from aptamer/quantum dot bioconjugates. Nanotechnology, 20 (17), 175503. Sassolas, A., Blum, L.J., and Leca-Bouvier, B.D. (2010) Homogeneous assays using aptamers. Analyst, 136 (2), 257–274. Khati, M. (2010) The future of aptamers in medicine. Journal of Clinical Pathology, 63 (6), 480–487. Li, J., Tan, S., Chen, X., Zhang, C.Y., and Zhang, Y. (2011) Peptide aptamers with biological and therapeutic applications. Current Medicinal Chemistry, 18 (27), 4215–4222. Song, K.M., Lee, S., and Ban, C. (2012) Aptamers and their biological applications. Sensors, 12 (1), 612–631. Li, Y., Guo, L., Zhang, Z., Tang, J., and Xie, J. (2008) Recent advances of aptamer sensors. Science in China. Series B, Chemistry, Life Sciences & Earth Sciences, 51 (3), 193–204. Hamaguchi, N., Ellington, A., and Stanton, M. (2001) Aptamer beacons for the direct detection of proteins. Analytical Biochemistry, 294 (2), 126–131. Ellington, A.D. and Szostak, J.W. (1990) In vitro selection of RNA molecules that bind specific ligands. Nature, 346 (6287), 818–822. Tuerk, C. and Gold, L. (1990) Systematic evolution of ligands by exponential enrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science, 249 (4968), 505–510. Bruno, J.G., Carrillo, M.P., and Phillips, T. (2008) Development of DNA aptamers to a foot-and-mouth disease peptide for competitive FRET-based detection. Journal j319 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine 245 246 247 248 249 250 251 252 253 254 of Biomolecular Techniques: JBT, 19 (2), 109–115. Chang, H., Tang, L., Wang, Y., Jiang, J., and Li, J. (2010) Graphene fluorescence resonance energy transfer aptasensor for the thrombin detection. Analytical Chemistry, 82 (6), 2341–2346. Juskowiak, B. (2006) Analytical potential of the quadruplex DNA-based FRET probes. Analytica Chimica Acta, 568 (1–2), 171–180. Li, J.W.J., Fang, X.H., and Tan, W.H. (2002) Molecular aptamer beacons for real-time protein recognition. Biochemical and Biophysical Research Communications, 292 (1), 31–40. Nutiu, R. and Li, Y.F. (2003) Structureswitching signaling aptamers. Journal of the American Chemical Society, 125 (16), 4771–4778. Dong, H., Gao, W., Yan, F., Ji, H., and Ju, H. (2010) Fluorescence resonance energy transfer between quantum dots and graphene oxide for sensing biomolecules. Analytical Chemistry, 82 (13), 5511–5517. Wang, Y., Bao, L., Liu, Z., and Pang, D.W. (2011) Aptamer biosensor based on fluorescence resonance energy transfer from upconverting phosphors to carbon nanoparticles for thrombin detection in human plasma. Analytical Chemistry, 83 (21), 8130–8137. Heyduk, E. and Heyduk, T. (2005) Nucleic acid-based fluorescence sensors for detecting proteins. Analytical Chemistry, 77 (4), 1147–1156. Lee, W., Obubuafo, A., Lee, Y.I., Davis, L. M., and Soper, S.A. (2010) Single-pair fluorescence resonance energy transfer (spFRET) for the high sensitivity analysis of low-abundance proteins using aptamers as molecular recognition elements. Journal of Fluorescence, 20 (1), 203–213. Bruno, J.G., Carrillo, M.P., Phillips, T., Vail, N.K., and Hanson, D. (2008) Competitive FRET-aptamer-based detection of methylphosphonic acid, a common nerve agent metabolite. Journal of Fluorescence, 18 (5), 867–876. Freeman, R., Bahshi, L., Finder, T., Gill, R., and Willner, I. (2009) Competitive analysis of saccharides or dopamine by 255 256 257 258 259 260 261 262 263 264 boronic acid-functionalized CdSe–ZnS quantum dots. Chemical Communications, 7, 764–766. Zhang, C.Y. and Johnson, L.W. (2009) Single quantum-dot-based aptameric nanosensor for cocaine. Analytical Chemistry, 81 (8), 3051–3055. Endoh, T., Shintani, R., Mie, M., Kobatake, E., Ohtsuki, T., and Sisido, M. (2009) Detection of bioactive small molecules by fluorescent resonance energy transfer (FRET) in RNA–protein conjugates. Bioconjugate Chemistry, 20 (12), 2242–2246. He, F., Yu, M.H., and Wang, S. (2009) Magnetically assisted fluorescence ratiometric assays for adenosine deaminase using water-soluble conjugated polymers. Chinese Science Bulletin, 54 (8), 1340–1344. Cheng, A.K., Su, H., Wang, A., and Yu, H. Z. (2009) Aptamer-based detection of epithelial tumor marker mucin 1 with quantum dot-based fluorescence readout. Analytical Chemistry, 81 (15), 6130–6139. Vicens, M.C., Sen, A., Vanderlaan, A., Drake, T.J., and Tan, W.H. (2005) Investigation of molecular beacon aptamer-based bioassay for plateletderived growth factor detection. ChemBioChem, 6 (5), 900–907. Li, W., Yang, X., Wang, K., Tan, W., Li, H., and Ma, C. (2008) FRET-based aptamer probe for rapid angiogenin detection. Talanta, 75 (3), 770–774. Zhang, J., Wang, L., Zhang, H., Boey, F., Song, S., and Fan, C. (2010) Aptamerbased multicolor fluorescent gold nanoprobes for multiplex detection in homogeneous solution. Small, 6 (2), 201– 204. Verch, T. and Bakhtiar, R. (2012) Miniaturized immunoassays: moving beyond the microplate. Bioanalysis, 4 (2), 177–188. Ligler, F.S. (2009) Perspective on optical biosensors and integrated sensor systems. Analytical Chemistry, 81 (2), 519–526. Rios, A., Zougagh, M., and Avila, M. (2012) Miniaturization through lab-on-achip: utopia or reality for routine laboratories? A review. Analytica Chimica Acta, 740, 1–11. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 320 265 Sun, S., Ossandon, M., Kostov, Y., and 266 267 268 269 270 271 272 273 274 Rasooly, A. (2009) Lab-on-a-chip for botulinum neurotoxin a (BoNT-A) activity analysis. Lab on a Chip, 9 (22), 3275–3281. Chen, L., Lee, S., Lee, M., Lim, C., Choo, J., Park, J.Y., Lee, S., Joo, S.W., Lee, K.H., and Choi, Y.W. (2008) DNA hybridization detection in a microfluidic channel using two fluorescently labeled nucleic acid probes. Biosensors & Bioelectronics, 23 (12), 1878–1882. Jung, J., Chen, L.X., Lee, S., Kim, S., Seong, G.H., Choo, J., Lee, E.K., Oh, C. H., and Lee, S. (2007) Fast and sensitive DNA analysis using changes in the FRET signals of molecular beacons in a PDMS microfluidic channel. Analytical and Bioanalytical Chemistry, 387 (8), 2609–2615. Tavares, A.J., Noor, M.O., Vannoy, C.H., Algar, W.R., and Krull, U.J. (2012) Onchip transduction of nucleic acid hybridization using spatial profiles of immobilized quantum dots and fluorescence resonance energy transfer. Analytical Chemistry, 84 (1), 312–319. Hsieh, A.T.-H., Pan, P.J.-H., and Lee, A.P. (2009) Rapid label-free DNA analysis in picoliter microfluidic droplets using FRET probes. Microfluidics and Nanofluidics, 6 (3), 391–401. Zhang, Y. and Ozdemir, P. (2009) Microfluidic DNA amplification: a review. Analytica Chimica Acta, 638 (2), 115–125. Wang, F. and Burns, M.A. (2009) Performance of nanoliter-sized dropletbased microfluidic PCR. Biomedical Microdevices, 11 (5), 1071–1080. Baker, M. (2012) Digital PCR hits its stride. Nature Methods, 9 (6), 541–544. Pekin, D., Skhiri, Y., Baret, J.C., Le Corre, D., Mazutis, L., Ben Salem, C., Millot, F., El Harrak, A., Hutchison, J., Larson, J.W., Link, D.R., Laurent-Puig, P., Griffiths, A. D., and Taly, V. (2011) Quantitative and sensitive detection of rare mutations using droplet-based microfluidics. Lab on a Chip, 11 (13), 2156–2166. Girkin, J.M., Mohammed, M.I., and Ellis, E. (2011) A miniaturized integrated biophotonic point-of-care genotyping system. Faraday Discussions, 149, 115–123. 275 Chang, C.C., Chen, C.C., Wei, S.C., Lu, H. 276 277 278 279 280 281 282 283 284 285 286 H., Liang, Y.H., and Lin, C.W. (2012) Diagnostic devices for isothermal nucleic acid amplification. Sensors, 12 (6), 8319–8337. Hoetzer, B., Medintz, I.L., and Hildebrandt, N. (2012) Fluorescence in nanobiotechnology: sophisticated fluorophores for novel applications. Small, 8 (15), 2297–2326. Liu, C.P., Cheng, S.H., Chen, N.T., and Lo, L.W. (2012) Intra/inter-particle energy transfer of luminescence nanocrystals for biomedical applications. Journal of Nanomaterials, 2012, Article# 706134, 1–9. Chi, X., Huang, D., Zhao, Z., Zhou, Z., Yin, Z., and Gao, J. (2012) Nanoprobes for in vitro diagnostics of cancer and infectious diseases. Biomaterials, 33 (1), 189–206. Algar, W., Susumu, K., Delehanty, J.B., and Medintz, I.L. (2011) Semiconductor quantum dots in bioanalysis: crossing the valley of death. Analytical Chemistry, 83 (23), 8826–8837. Boeneman, K., Delehanty, J.B., Susumu, K., Stewart, M.H., Deschamps, J.R., and Medintz, I.L. (2012) Quantum dots and fluorescent protein FRET-based biosensors. Advances in Experimental Medicine and Biology, 733, 63–74. Zhang, Y. and Wang, T.H. (2012) Quantum dot enabled molecular sensing and diagnostics. Theranostics, 2 (7), 631– 654. Kim, G.B. and Kim, Y.P. (2012) Analysis of protease activity using quantum dots and resonance energy transfer. Theranostics, 2 (2), 127–138. Jin, Z. and Hildebrandt, N. (2012) Semiconductor quantum dots for in vitro diagnostics and cellular imaging. Trends in Biotechnology, 30 (7), 394–403. Zhang, C.Y. and Hu, J. (2010) Single quantum dot-based nanosensor for multiple DNA detection. Analytical Chemistry, 82 (5), 1921–1927. Diamandis, E.P. (1988) Immunoassays with time-resolved fluorescence spectroscopy: principles and applications. Clinical Biochemistry, 21 (3), 139–150. Rantanen, T., Jarvenpaa, M.L., Vuojola, J., Arppe, R., Kuningas, K., and Soukka, T. j321 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License References j 7 In Vitro FRET Sensing, Diagnostics, and Personalized Medicine (2009) Upconverting phosphors in a dualparameter LRET-based hybridization assay. Analyst, 134 (8), 1713–1716. 287 Charbonniere, L.J. and Hildebrandt, N. (2008) Lanthanide complexes and quantum dots: a bright wedding for resonance energy transfer. European Journal of Inorganic Chemistry, 21, 3241–3251. 288 Morgner, F., Stufler, S., Geissler, D., Medintz, I.L., Algar, W., Susumu, K., Stewart, M.H., Blanco-Canosa, J.B., Dawson, P.E., and Hildebrandt, N. (2011) Terbium to quantum dot FRET bioconjugates for clinical diagnostics: influence of human plasma on optical and assembly properties. Sensors, 11 (10), 9667–9684. 289 Tian, J., Zhao, H., Liu, M., Chen, Y., and Quan, X. (2012) Detection of influenza A virus based on fluorescence resonance energy transfer from quantum dots to carbon nanotubes. Analytica Chimica Acta, 723, 83–87. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 322 8 Single-Molecule Applications Thomas Pons 8.1 Introduction Recent years have witnessed major progress in fluorescence microscopy instrumentation, raising its detection sensitivity to the single-molecule level. Observation of single molecules has since brought a wealth of information and allowed a better understanding of many physical, chemical, and biological processes [1–6]. Singlemolecule fluorescence has become in particular a powerful tool to study biomolecular functions [7–12]. Indeed, these functions most often involve conformational changes and/or multimolecular association/dissociation. This implies that, in the absence of an external synchronization, biomolecules in solution fluctuate between different states independently of each other. Whereas ensemble experiments provide only average measurements over all these different states, singlemolecule measurements can reveal both heterogeneity in the population (in the equilibrium distribution of states) and dynamics (i.e., sequence of transitions, frequencies, rates, etc.). Single-molecule F€orster (or Fluorescence) resonance energy transfer is certainly one of the most fertile single-molecule fluorescence techniques [13–19]. The vast majority of smFRET studies involve labeling of the target biomolecules with a donor and an acceptor fluorophore at specific sites. The FRET distance dependence translates changes in donor–acceptor separation distance into measurable photophysical parameters, such as donor/acceptor emission ratios, lifetimes, and anisotropy. This in turn provides a tool to follow conformational changes in a single biomolecule or association/dissociation dynamics in a single complex of interacting partners. Observation of single-molecule fluorescence signals raises several difficulties, including weak fluorescence signals and the need to isolate the signal from one molecule from the large background of other molecules, without perturbing the functional integrity of the biomolecule. These challenges can be overcome using mainly two categories of microscopy modalities. The first modality relies on immobilization of molecules on a substrate. A time trace of fluorescence signals FRET – Förster Resonance Energy Transfer: From Theory to Applications, First Edition. Edited by Igor Medintz and Niko Hildebrandt. Ó 2014 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2014 by Wiley-VCH Verlag GmbH & Co. KGaA. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License j323 j 8 Single-Molecule Applications can then be recorded, giving access to the full dynamics of molecular conformational changes. The second modality uses molecules dissolved in solution and freely diffusing in and out of a confocal detection volume, resulting in the rapid acquisition of series of short fluorescence bursts from many individual molecules. Analysis of these bursts reveals sample heterogeneity and allows identification of subpopulations and their repartition under equilibrium conditions. For each modality, we will briefly present the corresponding experimental techniques and analysis methods, discuss their capabilities, and present a few examples to illustrate their potential applications, with a strong focus on biophysical studies. Finally, we will present other single-molecule FRET schemes that involve multiple interacting FRET partners. 8.2 Single-Molecule FRET of Immobilized Molecules This section presents an overview of experimental techniques used to immobilize biomolecules on substrates, with the standard data analysis methods and some illustrative examples of applications. An advanced trace analysis technique is finally presented. 8.2.1 Experimental Setup 8.2.1.1 Molecule Immobilization Protocols used for molecule immobilization must be carefully designed to attach the biomolecule without interfering with its functionality and to avoid its nonspecific interactions with the substrate. One of the most common methods used for biomolecule immobilization is the biotinylation of the biomolecule and its subsequent attachment to surface-bound streptavidin [16]. This provides a highly specific interaction with a high affinity and allows subsequent washing of the surrounding solution without risking the detachment of the biomolecule of interest. However, extreme care must be taken to minimize potential interactions between the biomolecule and the rest of the substrate surface. Single DNA and RNA studies are usually performed using substrates coated with biotinylated BSA and then streptavidin [20,21]. Several studies have indeed verified that the conformation of oligonucleotides is not perturbed when immobilized on these surfaces [20,22,23], probably thanks to the electrostatic repulsion between the negatively charged oligonucleotides and the negatively charged glass, BSA, and streptavidin at neutral pH. In contrast, single proteins tend to present much more pronounced nonspecific adsorption on these surfaces and require better passivated substrates. Several methods have been proposed to reduce these undesired interactions. Most of them require the glass slide to be activated with aminosilanes first, followed by the covalent coupling of molecules to form a “furtive” coating, with a few biotin Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 324 groups for further specific biomolecule attachment. Covalently surface-linked and cross-linked BSA layers present a more uniform and robust surface coverage compared to their adsorbed counterparts, which result in a lower level of nonspecific protein adsorption [24]. Dense layers of polyethylene glycol are another popular coating and can indeed strongly reduce nonspecific adsorption when sufficiently long PEG chains are used [25]. These flexible PEG brushes have, however, been shown to intermingle with attached proteins and modify their conformation after a cycle of denaturation and refolding [26]. In comparison, cross-linked star-shaped PEG coatings seem to reduce these interactions since individual immobilized proteins refolded in their initial conformation after successive exposition to denaturation and refolding buffers [26,27]. Finally, an interesting alternative to direct attachment of biomolecules to the substrate is their confinement in lipidic nanovesicles [28]. The biomolecules are encapsulated within large (typically 100 nm) unilamellar vesicles containing a few biotinylated lipids, which allow subsequent immobilization of the vesicles on avidinfunctionalized supported lipid bilayers [28–31]. This method is attractive since the biomolecule remains in solution but is confined in a volume smaller than the diffraction limit of the microscope, and is therefore available for prolonged observation. The lipid vesicle is impermeable to the biomolecule of interest but may be made permeable to other smaller molecules by incorporating small pores in its membrane. This allows the controlled modification of the chemical environment of the biomolecule (ions, nucleotides, etc.) from the outside of the nanovesicle container [32–34]. Detailed sample preparation protocols are available in Ref. [33] and references therein. Irrespective of the immobilization technique used, appropriate control experiments must be performed to ensure that the biomolecules are indeed not perturbed. These may include testing different immobilization techniques, measuring their enzymatic activity, checking for dye anisotropy, and so on (see Section 8.2.3). 8.2.1.2 Fluorophore Photostability The choice of fluorescent labels must be optimized to provide strong and stable signals. The fluorophores should possess high extinction coefficients, fluorescence quantum yields, and photostability to allow long observation times and high signalto-noise ratios, and present limited photophysical effects (e.g., transient blinking). The most common single-molecule fluorophore pairs include Cy3–Cy5 and their Atto and AlexaFluor equivalents. The excitation rate of dyes in single-molecule experiments is much higher compared to ensemble measurements, leading to much faster photobleaching. Several reactants should therefore be added to the buffer solution to enhance the dye photostability. Molecular oxygen should be removed since it accelerates photobleaching through the formation of radical species. This is usually obtained using an enzymatic oxygen scavenger system composed of glucose oxidase, catalase, and b-D-glucose [35]. Other antioxidants may also be used, such as propyl gallate, ascorbic acid (vitamin C) [36], and cysteamine [37]. Oxygen is, however, also an efficient quencher of the triplet state, and removing it increases the lifetime of this dark state. Other triplet state quenchers thus need to j325 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.2 Single-Molecule FRET of Immobilized Molecules j 8 Single-Molecule Applications be added in the buffer solution, such as Trolox [37], an analogue of vitamin E, or mercaptoethylamine [36]. The nature and optimal concentration of oxygen and radical scavengers and triplet state quenchers may depend on the type of dye used [36]. Alternatively, oxygen removal may be coupled with the addition of an appropriate combination of reducing and oxidizing agents, such as ascorbic acid and methylviologen [38]. 8.2.1.3 Optical Setup The solution above the surface is easily washed from residual fluorophore-labeled molecules; however, fluorescence from the immobilized smFRET molecules must still be isolated from out-of-focus background to improve the signal-to-noise ratio. This can be achieved by confocal microscopy or most frequently by total internal reflection fluorescence (TIRF) microscopy [9,39,40]. Confocal detection consists in focalizing a laser beam through an objective using a pair of excitation and detection pinholes to eliminate fluorescence photons from outside a three-dimensional, diffraction-limited (<1 mm3) confocal volume. The fluorescence signals from molecules inside this volume are detected using photomultipliers or avalanche photodiodes, allowing high acquisition speed. However, these detectors only acquire data from one molecule at a time, and measuring a statistically significant number of single molecules becomes very time consuming. In contrast, TIRF microscopy uses wide-field illumination with an excitation beam that hits the substrate with an angle larger than the total internal reflection angle [40]. No light thus propagates into the medium above but the excitation light is confined to a thin (<200 nm) evanescent layer above the surface, and the intensity decays exponentially from the surface. Under these conditions, excitation is confined to fluorophores on or immediately above the substrate, and eliminates background from the solution above. In practice, this may be realized by focusing a laser beam on the edge of a high-NA objective back pupil to create a parallel beam tilted with respect to the objective optical axis. In this case, fluorescence photons are collected through the same objective. Alternatively, the laser may be focused on a prism placed on top of a thick quartz slide to create the evanescent wave (see Figure 8.1), and a high NA objective is placed below the sample to collect fluorescence photons. Fluorescence light is then split into donor and acceptor channels using dichroic and bandpass filters to form two separate images on a high-sensitivity, low-noise cooled EM-CCD camera. This allows parallel imaging of typically up to a few tens of single FRET pairs. 8.2.2 Data Analysis The first step in the analysis of single-molecule FRET data obtained by wide-field TIRF microscopy is the isolation of pixels and group of pixels containing single donor–acceptor pair signals. This is usually performed by selecting pixels above a predefined threshold, averaging a small area (e.g., 5 5 or 7 7 pixels) around the center pixel to integrate the point spread function of the microscope, and removing an average background value corresponding to neighboring “empty” pixels. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 326 Figure 8.1 (a) SmFRET setup for the observation of molecules immobilized on a substrate, using prism-based (i) or objectivebased (ii) TIRF microscopy. The donor and acceptor fluorescence images are separated into two halves. (b) Example of smFRET donor and acceptor and FRET time traces showing acceptor blinking and photobleaching events and three distinct FRET states. (Reproduced with permission from Ref. [16]. Copyright 2008, Macmillan Publishers Ltd.) Molecules with only one active fluorophore present donor-only or acceptor-only fluorescence signals and may then be discarded. The time traces obtained from signal intensities in the donor and acceptor channels, SD and SA, and their ratio, Eapp ¼ SA/(SA þ SD), qualitatively reflect the evolution of the separation distance between the two fluorophores with time. Low Eapp values correspond to long j327 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.2 Single-Molecule FRET of Immobilized Molecules j 8 Single-Molecule Applications separation distances, while high Eapp values correspond to shorter distances. There are, however, several factors that should be taken into account for a more quantitative analysis and calculation of true FRET efficiencies E, including direct acceptor excitation, spectral cross-talks of the donor emission into the acceptor channel and vice versa, and differences in fluorescence quantum yields and collection efficiencies between the two fluorophores. The signals in the donor and acceptor channels are related to the real molecular fluorescence intensities IA and ID and to the collection efficiencies through SD ¼ RDD I D þ RAD I A þ BA ; ð8:1Þ SA ¼ RAA IA þ RDA I D þ BD ; ð8:2Þ where Rij describes the instrument response and collection efficiency of signal i into channel j (i, j ¼ D, A) and BA,D represents the background signal from the detector (dark signal) and from spurious photons. The Rij elements may be determined by careful calibration using samples composed of donor-only and acceptor-only single fluorophores or fluorescent beads to correctly evaluate the corresponding ID and IA intensities. Due to the spectral shape of organic fluorophores, usually only the donor signal leaks into the acceptor channel, and the signals may be corrected by ID ¼ aD SD BD ; ð8:3Þ IA ¼ aA SA BA bSD ; ð8:4Þ where a and b constants take into account effects of channel cross-talks and detection efficiencies. Fluorescence intensities of the donor and acceptor molecules depend on their respective absorption cross sections, fluorescence quantum yield, and FRET efficiency as follows: ID ¼ sD I exc ð1 EÞWD ; ð8:5Þ IA ¼ ðsD E þ sA ÞI exc WA ; ð8:6Þ where s D(A) and WD(A) are the donor (acceptor) excitation cross section at the laser excitation wavelength and fluorescence quantum yield, respectively, Iexc is the laser excitation intensity, and E is the FRET efficiency. Assuming that the acceptor direct excitation is negligible (s D E s A ) or correctly accounted for, the FRET efficiency can then be evaluated as E ¼ I A =ðI A þ cI D Þ; c¼ WA : WD ð8:7Þ ð8:8Þ However, this assumes that c is identical for all individual FRET pairs. This is not necessarily true because of inhomogeneity in the fluorophore environment due to the substrate or the macromolecule conformation. However, the values of the c correction factor and the FRET efficiency may often be determined for each individual FRET pair. Indeed, the limited donor and acceptor photostability finally leads to permanent photobleaching of both emitters. In the case of the common Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 328 Cy3–Cy5 pair, the Cy5 acceptor often photobleaches before the Cy3 donor does. This leads to an instantaneous suppression of FRET processes and a corresponding recovery of donor emission. The amplitude of the donor recovery is directly related to the FRET efficiency before photobleaching since E ¼ (ID0 ID)/ID0, where ID0 is the final donor intensity after photobleaching, in the absence of any acceptor. In addition, the c factor corresponds to the ratio of the intensity changes before and after photobleaching DID and DIA: c ¼ DIA/DID. It can therefore be useful to compare the c values obtained for each individual molecule to the value obtained by average measurements. Single dyes often display complex photophysics that need to be taken into account for a correct single-molecule FRET interpretation. For example, organic dyes tend to “blink” and present short dark periods attributed to a triplet state (Figure 8.1b) [41–43]. Moreover, additional donor dye–acceptor dye interactions may take place at short separation distances [44]. These effects should not be confused with abrupt changes of FRET efficiencies. A good safeguard against incorrect interpretation of smFRET data is to look at the weighted sum of the fluorescence signals c ID þ IA, which should remain constant when FRET is the sole source of fluorescence fluctuations. Analysis of single-molecule FRET trajectories usually starts with identifying the different FRET states presented by the observed biomolecules. This is most often performed using simple thresholding, that is, defining a specific FRET efficiency range for each state (e.g., 0 < E < 0.2 for state s1, 0.35 < E < 0. 5 for state s2, and so on; see Figure 8.1). The first available information is the average distribution of FRET values (often presented as a histogram) or of states (e.g., at any given time molecules have a P1 ¼ 50% probability of being in state s1, P2 ¼ 20% probability of being in state s2, and so on). The free energy Gi of each conformation may then be simply evaluated using Gi ¼ kT ln(Pi) [45]. The second important information available is the sequence of conformational changes, for example, determining whether transitions always occur from s1 to s2, and then to s3, or directly from s1 to s3, and the frequency of the different transitions. Finally, another important parameter is the distribution of dwell time of each state, as this can be directly related to the corresponding transition rates. In particular, for states involved in single-rate kinetics, the distribution of dwell times t may be fitted with a monoexponential decay, exp(kt), where k is the transition rate out of the state. In the following section, we will present a few typical examples to illustrate the potential and limitations of smFRET on immobilized molecules. 8.2.3 Applications Since its demonstration using near-field scanning microscopy [46] in 1996, soon followed by its application to confocal microscopy [41,47], single-pair FRET measurements have been applied to a wide range of immobilized molecules, including oligonucleotides and proteins, to study various problems such as folding kinetics and bimolecular interactions. j329 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.2 Single-Molecule FRET of Immobilized Molecules j 8 Single-Molecule Applications The hairpin ribozyme has been extensively studied by smFRET and provides a good illustration of what can be achieved using smFRET on immobilized molecules: identifying subpopulations, measuring transition rates between these subpopulations, and the sequence of events occurring during a particular reaction. This ribozyme is an RNA enzyme capable of cleaving a specific RNA substrate into two products. The proposed reaction pathway starts with binding of the substrate RNA to the ribozyme. The hairpin ribozyme then consists in a four-arm DNA junction containing two internal loops. The ribozyme fluctuates between an extended unfolded conformation, in which the loops are far apart, and an active folded conformation, in which the two internal loops are interacting and close to each other. In the folded state, the cleavage reaction occurs and the resulting products finally dissociate from the ribozyme [48]. These conformational fluctuations may thus be observed by labeling one loop with a donor fluorophore and the other loop with an acceptor, and measuring changes in the FRET interactions on single molecules. In 2002, smFRET experiments were performed on a minimal form of the hairpin. Several FRETstates were identified with high FRET values corresponding to “folded” conformations and lower FRET values to “unfolded” conformations [49]. The cleavage reaction of the substrate finally occurs leading to a third distinct “cleaved” state. It was shown in particular that 90–95% of the single ribozyme molecules jumped to the cleaved state from the folded state, not from the unfolded state. The remaining 5–10% of FRET trajectories were attributed to contributions of short-lived folded state that were not observable with the limited (2 s) time resolution. SmFRET measurements thus allowed verification of the proposed reaction pathway, with cleavage occurring only in the folded state. In addition, measuring the dwell times in both folded and unfolded states revealed simple rate kinetics for the folding transition, but more complex multiexponential dynamics for unfolding transitions. This suggested the existence of different folded states. Moreover, a large heterogeneity in the unfolding kinetics was observed, with some time traces showing predominantly fast unfolding and some other slow unfolding. This conformational heterogeneity was confirmed in subsequent observations by Tan in 2003 on the natural form of the hairpin [50]. These experiments also revealed that the system did not show a single unfolded but rather two rapidly interconverting unfolded states. Transitions to the folded state occurred from the unfolded states possessing the higher FRET efficiency, therefore called a “proximal” state. The dynamics of conformational fluctuations depended strongly on Mg2þ concentrations. When the transitions were slow enough, it was possible to construct a histogram of dwell times of the molecule in the proximal state (Figure 8.2a). This histogram could be fitted with a monoexponential decay, with a decay time characteristic of the transition between the proximal and the distal states (Figure 8.2b). At low Mg2þ concentrations, however, the fluctuations became too fast to be clearly directly measured from single-molecule traces (Figure 8.2c). Fast smFRET dynamics may be more clearly resolved by cross-correlating the donor and acceptor fluorescence intensity traces. Indeed, changes in FRET efficiency result in anticorrelated intensity variations of the donor and acceptor fluorophores. The crosscorrelation function could then be fitted in turn with a monoexponential decay. In Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 330 Figure 8.2 (a) SmFRET time trace showing transitions between folded (F), proximal (Up), and distal (Ud) states. When transitions are clearly resolved, the histogram of dwell times in state Up reveals the transition time constant (here, 96 ms). (b–d) When transitions are faster, cross-correlation of donor and acceptor traces allows access to short transition time constants, down to 1 ms. (Reproduced with permission from Ref. [50]. Copyright 2003, National Academy of Sciences, USA.) this study, this analysis yielded transition rates of up to 1000s1 between the two unfolded states (Figure 8.2d). Finally, cleavage reaction kinetics were examined on a single hairpin basis from smFRET measurements and also showed a high heterogeneity. SmFRET data analysis must provide strong evidence that the molecule immobilization on the substrate does not modify its conformational dynamics. To test whether the previously observed heterogeneity could originate from nonspecific interactions between the immobilized hairpin and the substrate, these experiments were reproduced on single hairpins encapsulated inside 100–200 nm lipid vesicles [30]. This conformation is very different since the RNA is confined near the surface but not attached to it. Again, 50-fold variations in the dwell times of folded and unfolded states were observed. This confirmed that folding heterogeneity was indeed intrinsic to the hairpin and possibly attributable to heterogeneity in the conformation of the loop substructures, not to nonspecific interactions with the surface. While oligonucleotides present in general low levels of nonspecific interactions with the substrate, possibly due to repulsive electrostatic interactions, proteins are much more complex macromolecules that interact more strongly with the substrate and its functionalization layer. For example, Rhoades had studied the folding fluctuations of vesicle-encapsulated adenylate kinase (AK), a 214-amino acid protein j331 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.2 Single-Molecule FRET of Immobilized Molecules j 8 Single-Molecule Applications [29]. To test whether AK proteins were interacting nonspecifically with the vesicle walls, he first observed the fluorescence polarization values of AK labeled with the donor fluorophore only. All observed molecules showed very low polarization values, due to the fast rotational motion of the fluorophore that randomizes the polarization of consecutive fluorescence photons. On the contrary, proteins adsorbed on glass showed a much broader polarization distribution due to immobilization of the protein. In addition, he found that the distribution of FRET values was significantly different between vesicle-encapsulated and surface-immobilized proteins, suggesting a partial denaturation of the protein when adsorbed on the glass substrate and emphasizing the importance of the immobilization strategy. Most of the observed time traces showed single FRET values due to the folding transition being slower than the average fluorophore lifetime (10–20 s due to photobleaching). However, in smFRET traces showing at least one transition, histograms of FRET values showed mainly two subpopulations corresponding to folded and unfolded states. Analyzing the amplitude of changes in FRET values showed a large spread of the transitions, indicating a large heterogeneity of the folding reaction, and the existence of several intermediate folded states. In addition, the authors show that many molecules exhibit slow transitions (>1 s). These slow transitions were interpreted as continuous directed conformational changes, possibly slowed down by local traps. Intermolecular, not only intramolecular, interactions between DNA, RNA, and proteins may also be probed by smFRET to examine reaction pathways and kinetics. Ha et al. have, for example, examined DNA unwinding by the Rep helicase [21]. DNA probes were composed of an acceptor strand attached to a polymer-coated surface and a complementary donor strand. The two strands form a junction between singlestranded and double-stranded DNA (dsDNA). Unwinding of the dsDNA portion by helicases increased the separation between the dyes and thus reduced FRET interactions. Some time traces showed complete unwinding with a complete suppression of FRET due to fast diffusion of the donor strand away from the immobilized acceptor strand. However, some time traces show stalls in the DNA unwinding, with occasional rewinding. These transient events could not have been detected in ensemble experiments. Further experiments examining the kinetics of helicase binding and DNA unwinding under different concentration of Rep helicases were able to show that DNA unwinding must involve the interaction of more than one Rep protein. In the preceding example, only one of the interacting partners, the dsDNA, was labeled with fluorophores. This allowed tracking conformational changes of a few DNA molecules in the presence of a high concentration of Rep helicases (up to 100 nM) to increase the probability of Rep–DNA intermolecular interactions. However, situations where the molecules do not undergo drastic conformational changes upon interaction require labeling of both interacting molecules to follow their binding and unbinding kinetics. In that case, one is confronted with two seemingly incompatible constraints: the need for a low concentration to ensure detection of isolated FRET pairs (typically <0.1 nM) and the need for a high concentration to ensure efficient interaction (>1 mM depending on association– dissociation constants). Vesicle co-encapsulation provides a way to reconcile these Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 332 Figure 8.3 (a) Schematics showing two interacting proteins trapped inside a lipid nanovesicle. (b) Model for the protein interactions with one unbound and two distinct bound states. (c–e) Histograms of dwell times with the corresponding derived transition rates. (Adapted with permission from Ref. [31]. Copyright 2009, American Chemical Society.) two aspects [34]. The effective concentration inside the nanovesicle may be as high as tens of mM due to the confined space (1 molecule complex in a 100 nm sphere). Vesicles may, however, be immobilized on a substrate with a sufficient spatial separation to allow visualization of single vesicles. This experimental setup is thus particularly interesting to study weakly interacting molecules. Benitez et al., for example, have studied interactions between two copper-binding proteins, Hah1 and WDP, encapsulated in lipid nanovesicles (Figure 8.3a) [31]. Each protein was labeled at a C-terminal cysteine residue with a donor or acceptor fluorophore. The presence of only one protein pair was verified by examining photobleaching steps. FRET traces showed three distinct FRET states attributed to one unbound state and two different bound states. Transition rates (binding, dissociation, and conformational change) for each state were then derived from histograms of the different dwell times (Figure 8.3b). In another study, Cisse et al. developed methods to introduce nanometer-sized pores into the vesicle walls [32]. These pores were created either by incorporation of bacterial toxins or by bringing the vesicle at the lipid phase transition temperature, triggering lipid packing defects. These pores were impermeable to large molecules and maintained confinement of the studied DNA and protein molecules, but allowed modulating the concentration of ATP. The authors could then follow the interaction dynamics of the same interacting molecules under different ATP conditions. The change in the chemical environment indeed induced an observable change in the interaction between proteins and DNA. Finally, it should be noted that the vesicle lipid membrane constitutes a promising platform to study membrane-anchored proteins [34]. j333 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.2 Single-Molecule FRET of Immobilized Molecules j 8 Single-Molecule Applications 8.2.4 Analyzing Complex FRET Trajectories In most of the above examples, FRET trajectories were decomposed into two or three distinguishable FRET states and corresponding macromolecular conformations. This is easily achieved when the number of states is low and when FRET values can be clearly separated by a manually set threshold to discriminate between FRET states. This method is, however, ill adapted to situations with a higher number of states, or when it becomes difficult to discriminate the effects of noise or photophysical changes from a true conformational transition. These more complex FRET trajectories may be instead analyzed using hidden Markov modeling (HMM). This algorithm has been used in various applications such as speech recognition, cryptanalysis, and biophysics, including single-molecule FRET [51–53]. Its parameters consist of the emission probability functions and the transition probability matrix. The emission probability functions epA(B . . . )(g) describe the probability of a specific FRET ratio value g being detected when the system is in state A (B . . . ). Gaussian functions are generally used to include effects of real (conformational, photophysical) and noise-induced fluctuations. The transition probability matrix tp describes the probability of the system changing from a FRET state to another in the subsequent step and is directly related to the different transition monoexponential rates. The cumulated probability that a given FRET ratio sequence {g1, g2, g3} corresponds to a specific conformation trajectory {A ! A ! B} is then determined from the product of all corresponding emission and transition probabilities: epA(g1) epA(g2)epB(g3)tp(A ! A)tp(A ! B). Similar probabilities are calculated for all possible trajectories to determine the most probable trajectory. The Viterbi algorithm may help here to reduce computation costs [51,54]. In most experimental cases, neither the emission probability functions nor the transition matrix is known beforehand. These parameters are thus varied and optimized over a set of many different single-molecule FRET trajectories to obtain a maximized total probability. This finally yields the most probable emission probability functions for each state and the corresponding transition probability matrix. Several publicly available programs have been developed that allow application of HMM to single-molecule FRET data, such as HaMMy [51] (http://bio.physics.illinois.edu/HaMMy.html), QuB [52] (http://www.qub.buffalo.edu/wiki/index.php/Main_Page), and vb-FRET [55] (http://vbfret.sourceforge.net/). Hidden Markov modeling has been used, for example, by McKinney et al. to analyze binding of RecA proteins to single DNA molecules [51]. Since several RecA proteins can bind to a single DNA strand, different FRET states are observed, corresponding to different RecA:DNA ratios. While the high number of states precluded analysis using manually set thresholds between the different FRET states, HMM revealed the existence of five different states (0, 1, 2, 3, and 4 associated RecA proteins). Transition probabilities were high only between neighboring states, suggesting that RecA proteins bind and dissociate one by one to the DNA strand. Further analysis showed that association rates increased with RecA concentration, while dissociation rates were independent. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 334 In another report, Abelson et al. study the conformational changes of a small mRNA strand during in vitro splicing [56]. This complex multistep reaction involves an ensemble of small ribonucleoproteins and leads to the removal of specific intron sequences from the RNA strand. Figure 8.4a shows a typical fluorescence and FRET ratio time trace and the corresponding most probable sequence of FRET states. Figure 8.4b shows the transition density plot, representing the number of transitions observed from an initial FRET state (horizontal axis) to a particular final FRET state (vertical axis) in a collection of single-molecule FRET trajectories. This plot is used as a visual representation of the different observed transitions. Sometimes, however, a small number of molecules exhibit an unusually high number of fast transitions, which become therefore strongly emphasized in this representation, while slow transitions may be underrepresented. To avoid this problem, another representation may be used with population-weighted and kinetically indexed transition density (POKIT) plots. POKIT plots represent as concentric circles the fraction of molecules undergoing a specific transition at least once. This avoids giving too much importance to a small number of molecules exhibiting a large number of transitions. The average dwell time in an initial state before undergoing a specific transition is coded with different colors (e.g., red corresponding to fast transitions, green to slower ones, and so on). In the particular example shown in Figure 8.4c, these graphs summarize several important characteristics of the studied Figure 8.4 (a) Typical donor and acceptor fluorescence and FRET time traces in the splicing buffer, with the corresponding HMMderived traces; corresponding TDP (b) and POKIT (c) plots; (d) POKIT plot in ATP-depleted buffer. (Reproduced with permission from Ref. [56]. Copyright 2010, Macmillan Publishers Ltd.) j335 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.2 Single-Molecule FRET of Immobilized Molecules j 8 Single-Molecule Applications molecular system. It allows a quick evaluation of the number of distinct FRET states. HMM analysis of this system revealed the existence of up to 10 or 11 states but transitions occur predominantly between 5 distinct states. One can also immediately see that transitions occur mostly between neighboring FRET states, implying a stepby-step reaction. Finally, the graph is symmetric with respect to the diagonal, which means that most transitions work in both directions and that they are reversible. These graphs also facilitate comparisons of different reaction conditions. For example, Figure 8.4d shows the POKIT plot under ATP-depleted conditions: both the nature and the dynamics of the observed transitions are modified. 8.3 Single-Molecule FRET of Freely Diffusing Molecules This section presents the second family of single-molecule FRET techniques, consisting in the detection of freely diffusing molecules in solution. We present the experimental techniques, along with a few examples to illustrate what can be achieved with solution smFRET experiments. Finally, we present advanced techniques that take advantage of the large volume of data available with solution smFRET to perform more refined data analysis. 8.3.1 Experimental Setup Detecting single molecules freely diffusing in solution presents several advantages and drawbacks. It is both simpler and more robust than imaging immobilized molecules. This is due in part to the absence of any substrate surface susceptible to interfere with the conformation of the molecule of interest, which greatly simplifies sample preparation and reduces possible sample-to-sample variations. On the other hand, the need to isolate the fluorescence signal of an individual molecule from its neighbors puts strong constraints on the detection scheme and chromophore concentration. The detection volume must indeed be sufficiently small to contain at most one molecule at a time. This is usually realized by confocal or two-photon excited fluorescence microscopy. Typically, an excitation laser is focused by a high numerical aperture objective into the sample solution a few microns above a glass coverslip. Fluorescence photons are collected through the same objective and separated from the excitation beam by a dichroic mirror and emission filters. Donor and acceptor photons are then separated by a second dichroic mirror, passed through additional emission filters, and finally detected on avalanche photodiode detectors (APDs). Appropriate pinholes are placed in the excitation and detection path at the image conjugate planes to ensure confocality. All detected fluorescence photons originate from molecules located inside the small (few femtoliters) diffraction-limited confocal volume. Alternatively, the sample solution may be introduced in a small glass capillary and flowed through the confocal observation Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 336 volume. This capillary may be integrated into a microfluidic device to probe, for example, chemical reactions at different times after mixing or the effects of heating by changing the location of the detection spot along a microfluidic channel [57]. Finally, zero-mode waveguides present an interesting alternative to isolate fluorescence signals from single molecules [58,59]. This technique uses nanometer-scale apertures fabricated in thin metal films on a transparent substrate. Light intensity decays very rapidly at the entrance of the apertures, creating effective detection volumes that are three orders of magnitude smaller than diffraction-limited confocal volumes. In practice, this enables to probe more concentrated solutions but reduces the time spent by the molecule in the detection spot. In concentrated solutions, Brownian motion and flow induce fluctuations in the detected fluorescence signals, which may be analyzed using fluorescence correlation spectroscopy (FCS) [60,61]. However, when the observed solution is dilute enough (typically a few tens of pM for confocal detection schemes), most of the time the confined observation volume does not contain any fluorophore, and for brief periods it contains a single biomolecule. The resulting traces thus present long periods of background noise interrupted by fluorescence bursts. The duration of these bursts corresponds to the time necessary for the biomolecule to diffuse out of the observation volume. Their intensity depends on the intrinsic fluorescence parameters described in Equations 8.3 and 8.4 for the observation of single immobilized molecules (excitation cross sections, fluorescence quantum yields, FRET efficiency, etc.). APD signals are collected by counting boards and time traces are then registered, which can be binned using different time resolutions. Fluorescence bursts above a predefined threshold value are then selected from the background noise and corrected from detection efficiencies and spectral cross-talks as described in Equations 8.1 and 8.2. The most simple and straightforward analysis of singlemolecule FRET data consists in displaying histograms of the number or population fraction of bursts as a function of the emission ratio g ¼ IA/(IA þ cID), where IA and ID are the acceptor and donor burst intensities, respectively, and c ¼ WA/WD is the correction factor already described above (Figure 8.5). The emission ratio g is closely related to the FRET efficiency provided all corrections are properly performed (Equation 8.5). It is, however, difficult to ensure that each fluorescence burst is appropriately corrected, due to the possible heterogeneity in the environment and photophysical properties of each fluorophore pair, and the emission ratio is often referred to in practice as the apparent FRET efficiency, Eapp. 8.3.2 Applications The first ratiometric measurements of single freely diffusing biomolecule FRET were developed practically simultaneously to those on surface-bound molecules [63–65]. Immediate advantages of solution measurements were to eliminate possible artifacts due to nonspecific interactions with the surface and to rapidly measure a large number of events to enable statistical analysis. Solution smFRET was rapidly applied to identify subpopulations in a heterogeneous ensemble. In particular, j337 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.3 Single-Molecule FRET of Freely Diffusing Molecules j 8 Single-Molecule Applications 100 Counts 80 Sample Obj 60 40 20 0 320 Dichroic 1 Notch filter Dichroic 2 LP filter APDD APDA 321 322 323 324 325 Time (s) Population fraction Excitation laser 0.1 U 0.08 0.06 D F 0.04 0.02 0 -0.2 0 0.2 0.4 0.6 0.8 1 1.2 Emission ratio Figure 8.5 Typical solution smFRET optical setup and experimental fluorescence burst time traces. Bursts are selected (arrows) when the sum of their donor and acceptor photons is larger than the predefined threshold (dashed line). The donor and acceptor burst intensities are then analyzed to provide a FRET histogram, showing, for example, donor-only (D), unfolded (U), and folded (F) species. (Adapted with permission from Ref. [62]. Copyright 2006, American Chemical Society.) biomolecule labeling with donor and acceptor fluorophores is often incomplete. In addition, a portion of the acceptor dyes may be nonfluorescent and/or nonabsorbing due to photobleaching and blinking. This often leads to a nonnegligible population of donor-only biomolecules. These donors are not quenched by FRET, and this may lead to underestimate FRET efficiencies in ensemble measurements. Solution smFRET allows an easy identification of those donor-only molecules as a zerocentered peak in apparent FRET efficiency histograms. Deniz et al. demonstrated in 1999 that they were able to isolate this population and take into account only signals from dual-labeled DNA double strands for sufficiently high average Eapp (>0.4) [64]. In addition, they varied the distance between the donor and acceptor dyes and observed a progressive decrease in FRET efficiency for larger separation distances. They showed that they were able to separate two dsDNA populations corresponding to different interdye distances. The sequence with the longer separation distance, DNA17, contained an enzyme target sequence between the donor and acceptor fluorophores, while the other, DNA7, did not. Cleavage of the DNA strands by the enzyme led to the separation of the fluorophores in the DNA17 population. The authors showed that the DNA17 peak in the Eapp histogram indeed disappeared while the peak around zero, corresponding to isolated donor molecules, increased accordingly. The peak corresponding to the DNA7 sequence remained unchanged in the process. Protein folding studies have benefited from smFRET as a tool to separate folded from unfolded species. The cold shock protein, for example, has served as an Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 338 excellent model system for protein folding due to its simple structure and its twostate behavior [66,67]. Under normal conditions, FRET efficiency histograms of dual-labeled protein solutions show the usual donor-only peak around zero and a peak at high FRET efficiency, corresponding to the compact folded state. Under denaturating conditions, this peak disappears while a new peak arises at lower FRET efficiencies, corresponding to unfolded states (Figure 8.6a) [68]. In addition, this second peak shifts toward lower FRET efficiencies as the concentration of denaturant increases, indicating that the protein further unfolds. Revealing this type of complex behavior is a unique force of single-molecule studies and would have been very difficult to extract from FRET ensemble measurements. In this simple solution experiment, one observes the equilibrium of folded and unfolded populations, but does not have access to out-of-equilibrium conditions such as transient states or the underlying folding and unfolding dynamics. This may be achieved using microfluidic devices in which different solvents, proteins, and molecules may be flowed in channels and mixed where the channels merge under controlled conditions of flow. Locating the detection volume at different points of the output channel allows probing different times after mixing of the reactants, depending on the flow speed and the distance between the detection volume and the mixer. Lipman et al. used this Figure 8.6 (a) Histograms of measured FRET efficiencies at various denaturant concentrations for labeled cold shock protein. (Reproduced with permission from Ref. [68]. Copyright 2002, Macmillan Publishers Ltd.) (b) Varying the distance between the mixing point of the microfluidic channel and the laser confocal volume allows probing different times after mixing. FRET histograms obtained for different times after mixing. (Reproduced with permission from Ref. [57]. Copyright 2003, American Association for the Advancement of Science.) j339 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.3 Single-Molecule FRET of Freely Diffusing Molecules j 8 Single-Molecule Applications principle to probe cold shock protein folding kinetics by smFRET [57]. They induced an abrupt drop of denaturant concentration in the protein solution by dilution. They observed that the unfolded proteins rapidly switch to a more compact (higher FRET) unfolded state before slowly reaching the new equilibrium between folded and unfolded states (Figure 8.6b). This experiment thus gives access to transient species that are not strongly represented under equilibrium conditions. The ability to eliminate artifacts due to imperfect biomolecule labeling and measure FRET efficiencies with a good precision was used, for example, to revisit polyproline peptides as rigid spectroscopic rulers. These studies provide a good illustration of what information can be extracted from smFRET measurements in terms of subpopulations and dynamics, and of the possible caveats. Polyprolines adopt a type II helix in aqueous solution [69] and have been used as rigid spacers to verify the distance dependence of FRET in the range of 1–12 prolines per peptide, corresponding to a 20–45 A range [70]. Schuler examined polyprolines in a larger size range, from 6 to 40 proline residues using a FRET pair with a larger R0 distance (54 A instead of 35 A in the earlier work) [71]. He found that solution of longer peptides contained a nonnegligible fraction of donor-only molecules, which could lead to errors in ensemble measurements. However, the authors observed a significant discrepancy between the observed FRET efficiencies, Eapp, and those theoretically predicted assuming a rigid peptide conformation. Shorter peptides showed lower Eapp than predicted; this discrepancy has been attributed to failure of the point-dipole FRET model at short interdye distances and to the absence of fast orientational averaging, and consequently an error in the estimation of the k2 factor. This was supported by polarization measurements. Long peptides showed a low residual steady-state polarization value of 0.05, corresponding to a near total reorientation of the dipole between excitation and emission. However, for shorter peptides, the donor decay was significantly accelerated due to FRET and became comparable to the reorientation speed, as shown by the increase of steady-state polarization value to 0.11. Replacing the usual k2 ¼ 2/3 value corresponding to the rapid orientational averaging by an approximation of random, but static orientations, the authors found theoretical FRET efficiencies compatible with the observed values. Longer peptides, on the contrary, showed higher FRET efficiencies than expected. These longer peptides must thus adopt different conformations, bringing the two dyes closer to each other, and allow for higher FRET rates. Molecular dynamics showed that the timescale of these fluctuations (0.1–10 ns) was much shorter than the fluorescence burst duration (1 ms). In this regime, the effect of conformational fluctuations on the apparent FRET efficiency depends on the comparison between the dynamics of the peptide conformation, the rotational correlation time of the fluorophores, and their fluorescence lifetime. In this study, fluorescence and anisotropy decay measurements showed that the dye rotation (0.3 ns) was faster than both the fluorescence lifetime (1–10 ns) and the expected conformational fluctuations (0.1–10 ns). The obtained theoretical estimations of FRET efficiencies were in the same range as the experimental values, confirming that these peptides Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 340 indeed adopt flexible conformations and showing the importance of considering dynamics for extracting structural information. Another regime exists when conformational fluctuations change slowly compared to the duration of fluorescence burst, that is, the time spent by the biomolecule in the detection spot before diffusing away. In that case, each molecule presents a different, static conformation when passing through the confocal volume. This results in a broadening of the FRET histograms. The width of FRET histogram peaks thus contains information about dynamics of conformational changes. Its analysis is, however, complicated by additional contributions from shot noise and photophysical effects. The shot noise standard deviation s sn arises from the finite number of photons in each time bin and provides a fundamental statistical lower limit to the width of the histogram peaks. It is defined by EÞ=ðn s2sn ¼ Eð1 A þ nD Þ; ð8:9Þ is the average FRETefficiency and nA þ nD is the total number of photons in where E the donor and acceptor channel. A limit on this number is provided by considering the threshold value below which fluorescence bursts are rejected. Additional effects of dye molecular fluctuations are difficult to isolate and characterize. A reference molecule with the same fluorophores but no conformational fluctuations slower than the time bins considered may be used to estimate a standard width s 0. Any excess width, defined by s 2 s 20, may then be attributed to slow conformational fluctuations. Theoretical FRET histogram widths have been derived by Gopich and Szabo as a function of fluctuation dynamics and acquisition speed [72,73]. Schuler et al. used this method to examine the folding kinetics of cold shock protein. They analyzed FRET histograms obtained under different denaturing conditions, where the protein fluctuated between unfolded and folded configurations (Figure 8.6a). They then observed that the FRET peak widths were not much larger than those for a short rigid polyproline peptide [67,68]. This allowed them to put an upper limit of 30 ms to the protein reconfiguration time. This conformational dynamics can be used to obtain information about the free energy barrier for folding, as discussed in Refs [67,68]. Best et al. have further used the distributions in FRET histograms to obtain information about conformational flexibility of long polyprolines [74]. They have used a pulsed excitation source and recorded the arrival time of each photon after the corresponding excitation pulse. Interestingly, the authors were able to rule out possible contributions to the distribution widths from acceptor photobleaching, by observing that histograms constructed from the first and second halves of the time bins were identical. If photobleaching had any influence, histograms from the second half of the bins would have been shifted to lower FRET values. They also ruled out contributions from acceptor random blinking by examining the statistics of strings of consecutive donor or acceptor photons. If there is no blinking, the probability of observing a sequence of n consecutive donor photons varies as (1 Eapp)n. In contrast, blinking of the acceptor would give rise to a higher probability of observing long strings of donor photons. The distributions of j341 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.3 Single-Molecule FRET of Freely Diffusing Molecules j 8 Single-Molecule Applications Figure 8.7 SmFRET results for polyproline-20 in TFE (a) and water (b). The yellow dashed line indicates shot noise-limited width of the distribution. Insets: The donor fluorescence decays for donor photons from the subpopulations with corresponding colors in the efficiency histograms. (Reproduced with permission from Ref. [74]. Copyright 2007, National Academy of Sciences, USA.) consecutive sequences of donor photons perfectly matched theoretical predictions in the absence of blinking. The ability to record the arrival times of photons after each excitation pulse allowed them to reconstruct fluorescence decay curves from selected molecular subpopulations [74,75]. They were able to show that, for polyproline peptides in water, photons from molecules showing up in the higher side of the FRET histogram peak yielded a faster decay curve compared to photons from molecules showing lower FRET values (Figure 8.7). This indicated that FRET value heterogeneity indeed originated from differences in FRET rates, that is, to (slow) conformational fluctuations, not from noise. In contrast, fluorescence decay curves were identical throughout the smFRET histogram measured in trifluoroethanol (TFE), indicating that conformational fluctuations were smaller in this solvent compared to water. The authors were further able to use fluorescence decay data as an additional tool to evaluate FRET efficiencies. To construct the fluorescence decay of donor-only species, they used photons from bursts showing a near-zero emission ratio. They then calculated from molecular simulations the predicted fluorescence decay curves taking into account FRET processes corresponding to different peptide conformations. This analysis, together with NMR measurements, allowed them to identify different peptide conformations in TFE and water due to isomerization of proline residues. Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 342 8.3.3 Advanced Solution smFRET Methods Single-molecule FRET methods described above are powerful tools to study heterogeneous populations but suffer from several limitations, due to sample preparation and complex fluorophore photophysics. For example, smFRET cannot distinguish between donor-only species and dual-labeled species with low FRET efficiencies: in both cases, only the donor fluorophore is detected. More generally, it cannot detect acceptor-only species, which limits studies of the interactions of a donor-labeled molecule A with an acceptor-labeled molecule B. To overcome these limitations, several techniques were introduced, based on alternate laser excitation (ALEX) and multiparameter fluorescence detection (MFD). These schemes may also be applied to surface-immobilized molecules, but are more adapted to solution diffusing probes, thanks to the large volume of data accessible in solution measurements. 8.3.3.1 Alternate Laser Excitation The ALEX method has been developed by Kapanidis et al. as an extension of smFRET, in which the excitation consists of interleaved pulses of two different wavelengths: one for donor excitation and the other for acceptor excitation [76,77]. The excitation wavelength switches more rapidly than the average fluorescence burst duration, so that many excitation pulses are used for each single diffusing molecule. Photon detection is synchronized with the excitation pattern providing four sets of data after cross-talk and background corrections. Two of these fluorescence time Aem traces correspond to standard smFRET measurements, denoted F Dem Dexc and F Dexc , respectively, for the donor and acceptor emissions under excitation of the donor. The other fluorescence time traces correspond to the emission of the donor and the Aem Dem acceptor under direct acceptor excitation, F Dem Aexc and F Aexc (in practice, F Aexc 0). The first two signals provide the usual apparent FRET efficiency, Eapp, using Equation 8.7. In addition, the sum of all donor and acceptor emissions under Aem donor excitation,F Dexc ¼ cF Dem Dexc þ F Dexc , and that under acceptor excitation, Aem F Aexc ¼ cF Dem Aexc þ F Aexc , are calculated. One can now also define a stoichiometry ratio, S, as S¼ F Dexc : F Dexc þ F Aexc ð8:10Þ This ratio is independent of FRET efficiency, since F Dexc sums all photons emitted after donor excitation and is corrected for the difference in quantum yields and detection efficiency between the two fluorophores. The laser excitation intensities are usually adjusted to yield F Dexc F Aexc . In that case, donor-only species display S 1, acceptor-only species display S 0, and dual-labeled molecules assume intermediate values. Kapanidis et al. demonstrated that this method enabled separation of subpopulations based on apparent FRET efficiency and stoichiometry [77]. They were, for example, able to correctly take into account acceptor-only species, and also the presence of dimers (e.g., two donor–one acceptor macromolecules). j343 Downloaded from https://onlinelibrary.wiley.com/doi/ by University Of Texas Libraries, Wiley Online Library on [13/09/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License 8.3 Single-Molecule FRET of Freely Diffusing Molecules j 8 Single-Molecule Applications Figure 8.8 Typical example of an ALEX smFRET experiment, with two dsDNA strands with different interdye distances. The S versus E plot reveals donor-only species (upper left), acceptor-only species (bottom right), and differe