FRET-based dynamic structural biology: Challenges, perspectives and an appeal for open-science practices

  1. Eitan Lerner  Is a corresponding author
  2. Anders Barth  Is a corresponding author
  3. Jelle Hendrix  Is a corresponding author
  4. Benjamin Ambrose
  5. Victoria Birkedal
  6. Scott C Blanchard
  7. Richard Börner
  8. Hoi Sung Chung
  9. Thorben Cordes
  10. Timothy D Craggs
  11. Ashok A Deniz
  12. Jiajie Diao
  13. Jingyi Fei
  14. Ruben L Gonzalez
  15. Irina V Gopich
  16. Taekjip Ha
  17. Christian A Hanke
  18. Gilad Haran
  19. Nikos S Hatzakis
  20. Sungchul Hohng
  21. Seok-Cheol Hong
  22. Thorsten Hugel
  23. Antonino Ingargiola
  24. Chirlmin Joo
  25. Achillefs N Kapanidis
  26. Harold D Kim
  27. Ted Laurence
  28. Nam Ki Lee
  29. Tae-Hee Lee
  30. Edward A Lemke
  31. Emmanuel Margeat
  32. Jens Michaelis
  33. Xavier Michalet
  34. Sua Myong
  35. Daniel Nettels
  36. Thomas-Otavio Peulen
  37. Evelyn Ploetz
  38. Yair Razvag
  39. Nicole C Robb
  40. Benjamin Schuler
  41. Hamid Soleimaninejad
  42. Chun Tang
  43. Reza Vafabakhsh
  44. Don C Lamb  Is a corresponding author
  45. Claus AM Seidel  Is a corresponding author
  46. Shimon Weiss  Is a corresponding author
  1. Department of Biological Chemistry, The Alexander Silberman Institute of Life Sciences, and The Center for Nanoscience and Nanotechnology, Faculty of Mathematics & Science, The Edmond J. Safra Campus, The Hebrew University of Jerusalem, Israel
  2. Lehrstuhl für Molekulare Physikalische Chemie, Heinrich-Heine-Universität, Germany
  3. Dynamic Bioimaging Lab, Advanced Optical Microscopy Centre and Biomedical Research Institute (BIOMED), Hasselt University, Belgium
  4. Department of Chemistry, University of Sheffield, United Kingdom
  5. Department of Chemistry and iNANO center, Aarhus University, Denmark
  6. Department of Structural Biology, St. Jude Children's Research Hospital, United States
  7. Laserinstitut HS Mittweida, University of Applied Science Mittweida, Germany
  8. Laboratory of Chemical Physics, National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health, United States
  9. Physical and Synthetic Biology, Faculty of Biology, Ludwig-Maximilians-Universität München, Germany
  10. Department of Integrative Structural and Computational Biology, The Scripps Research Institute, United States
  11. Department of Cancer Biology, University of Cincinnati School of Medicine, United States
  12. Department of Biochemistry and Molecular Biology and The Institute for Biophysical Dynamics, University of Chicago, United States
  13. Department of Chemistry, Columbia University, United States
  14. Department of Biophysics and Biophysical Chemistry, Department of Biomedical Engineering, Johns Hopkins University School of Medicine, Howard Hughes Medical Institute, United States
  15. Department of Chemical and Biological Physics, Weizmann Institute of Science, Israel
  16. Department of Chemistry & Nanoscience Centre, University of Copenhagen, Denmark
  17. Denmark Novo Nordisk Foundation Centre for Protein Research, Faculty of Health and Medical Sciences, University of Copenhagen, Denmark
  18. Department of Physics and Astronomy, and Institute of Applied Physics, Seoul National University, Republic of Korea
  19. Center for Molecular Spectroscopy and Dynamics, Institute for Basic Science and Department of Physics, Korea University, Republic of Korea
  20. Institute of Physical Chemistry and Signalling Research Centres BIOSS and CIBSS, University of Freiburg, Germany
  21. Department of Chemistry and Biochemistry, and Department of Physiology, University of California, Los Angeles, United States
  22. Department of BioNanoScience, Kavli Institute of Nanoscience, Delft University of Technology, Netherlands
  23. Biological Physics Research Group, Clarendon Laboratory, Department of Physics, University of Oxford, United Kingdom
  24. School of Physics, Georgia Institute of Technology, United States
  25. Physical and Life Sciences Directorate, Lawrence Livermore National Laboratory, United States
  26. School of Chemistry, Seoul National University, Republic of Korea
  27. Department of Chemistry, Pennsylvania State University, United States
  28. Departments of Biology and Chemistry, Johannes Gutenberg University, Germany
  29. Institute of Molecular Biology (IMB), Germany
  30. Centre de Biologie Structurale (CBS), CNRS, INSERM, Universitié de Montpellier, France
  31. Institüt of Biophysics, Ulm University, Germany
  32. Department of Biophysics, Johns Hopkins University, United States
  33. Department of Biochemistry and Department of Physics, University of Zurich, Switzerland
  34. Department of Bioengineering and Therapeutic Sciences, University of California, San Francisco, United States
  35. Physical Chemistry, Department of Chemistry, Center for Nanoscience (CeNS), Center for Integrated Protein Science Munich (CIPSM) and Nanosystems Initiative Munich (NIM), Ludwig-Maximilians-Universität, Germany
  36. Warwick Medical School, University of Warwick, United Kingdom
  37. Biological Optical Microscopy Platform (BOMP), University of Melbourne, Australia
  38. College of Chemistry and Molecular Engineering, PKU-Tsinghua Center for Life Sciences, Beijing National Laboratory for Molecular Sciences, Peking University, China
  39. Department of Molecular Biosciences, Northwestern University, United States
  40. Department of Physiology, CaliforniaNanoSystems Institute, University of California, Los Angeles, United States

Abstract

Single-molecule FRET (smFRET) has become a mainstream technique for studying biomolecular structural dynamics. The rapid and wide adoption of smFRET experiments by an ever-increasing number of groups has generated significant progress in sample preparation, measurement procedures, data analysis, algorithms and documentation. Several labs that employ smFRET approaches have joined forces to inform the smFRET community about streamlining how to perform experiments and analyze results for obtaining quantitative information on biomolecular structure and dynamics. The recent efforts include blind tests to assess the accuracy and the precision of smFRET experiments among different labs using various procedures. These multi-lab studies have led to the development of smFRET procedures and documentation, which are important when submitting entries into the archiving system for integrative structure models, PDB-Dev. This position paper describes the current ‘state of the art’ from different perspectives, points to unresolved methodological issues for quantitative structural studies, provides a set of ‘soft recommendations’ about which an emerging consensus exists, and lists openly available resources for newcomers and seasoned practitioners. To make further progress, we strongly encourage ‘open science’ practices.

Introduction

Understanding how biomolecules couple structural dynamics with function is at the heart of several disciplines and remains an outstanding goal in biology. Linking conformational states and their transitions to biochemical function requires the ability to precisely resolve the structure and dynamics of a biological system, which is often altered upon ligand binding or influenced by the chemical and physical properties of its environment. The most well-established structural biology tools have provided high-resolution ‘snapshots’ of states in a crystallized or frozen form (e.g., X-ray crystallography and single-particle cryo-electron microscopy, cryoEM) or an ensemble average of all contributing conformations (e.g., nuclear magnetic resonance, NMR; small-angle X-ray scattering, SAXS; small-angle neutron scattering, SANS; double electron-electron resonance, DEER; cross-linking mass spectrometry, XL-MS; ensemble-FRET). In recent years, further developments have enabled these conventional structural tools to detect conformational dynamics and reaction intermediates. For example, NMR techniques (Anthis and Clore, 2015; Clore and Iwahara, 2009; Palmer, 2004; Ravera et al., 2014; Sekhar and Kay, 2019) and electron paramagnetic resonance techniques (Jeschke, 2018; Jeschke, 2012; Krstić et al., 2011) have been advanced to study conformational dynamics and capture transient intermediates. Time-resolved crystallographic investigations have been employed to resolve functionally relevant structural displacements associated with a biological function (Kupitz et al., 2014; Moffat, 2001; Schlichting et al., 1990; Schlichting and Chu, 2000; Schotte et al., 2003). Advances in microfluidic mixing and spraying devices have enabled time-resolved cryoEM (Feng et al., 2017; Kaledhonkar et al., 2018) and cross-linking mass spectrometry (XL-MS or CL-MS) (Braitbard et al., 2019; Brodie et al., 2019; Chen et al., 2020; Iacobucci et al., 2019; Murakami et al., 2013; Slavin and Kalisman, 2018). Progress in computational methods has also afforded novel tools for examining biomolecular structure and dynamics. Each of these advances highlights an increased awareness that one needs to directly and continuously track the dynamical properties of individual biomolecules in order to understand their function and regulation.

In this context, FRET (referred to as fluorescence resonance energy transfer or Förster resonance energy transfer [Braslavsky et al., 2008]) studies at the ensemble and single-molecule levels have emerged as important tools for measuring structural dynamics over at least 12 orders of magnitude in time and mapping the conformational and functional heterogeneities of biomolecules under ambient conditions. FRET studies probing fluorescence decays at the ensemble level (Grinvald et al., 1972; Haas et al., 1975; Haas and Steinberg, 1984; Hochstrasser et al., 1992) (time-resolved FRET) permitted already in the early 1970s the study of structural heterogeneities on timescales longer than the fluorescence lifetime (a few ns). This approach is still used nowadays (Becker, 2019; Orevi et al., 2014; Peulen et al., 2017) and has been transferred to single-molecule studies. The ability to measure FRET in single molecules (Deniz et al., 1999; Ha et al., 1996; Lerner et al., 2018a) has made the method even more appealing. The single-molecule FRET (smFRET) approach has been extensively used to study conformational dynamics and biomolecular interactions under steady-state conditions (Dupuis et al., 2014; Larsen et al., 2019; Lerner et al., 2018a; Lipman et al., 2003; Margittai et al., 2003; Mazal and Haran, 2019; Michalet et al., 2006; Orevi et al., 2014; Ray et al., 2019; Sasmal et al., 2016; Schuler et al., 2005; Schuler et al., 2002; Steiner et al., 2008; Zhuang et al., 2000). It is notable that, in many mechanistic studies, it suffices to use FRET for distinguishing different conformations and determining kinetic rates such that absolute FRET efficiencies and thereby distances do not need to be determined. However, the ability to measure accurate distances and kinetics with smFRET has led to its emergence as an important tool in this new era of ‘dynamic structural biology’ for mapping biomolecular heterogeneities and for measuring structural dynamics over a wide range of timescales (Lerner et al., 2018a; Mazal and Haran, 2019; Sanabria et al., 2020; Schuler and Hofmann, 2013; Weiss, 1999).

Single-molecule FRET (smFRET) approaches have many advantages as a structural biology method, including:

  • sensitivity to macro-molecular distances (2.5–10 nm),

  • the ability to resolve structural and dynamic heterogeneities,

  • high-quality measurements with low sample consumption of the molecules of interest (low concentrations and low volumes), as the sample is analyzed one molecule at a time,

  • determination of structural transitions in equilibrium, hence without the need for synchronization,

  • the ability to detect (very) rare events. Indeed, in biology, the most interesting molecules to study are often the sparse, functionally active ones amidst a sea of inactive molecules,

  • high sensitivity and specificity for labeled molecules. As only the labeled molecule uniquely contributes to the detected signal, these tracers can also be applied as FRET-reporters in crowded environments (Dupuis et al., 2014; Soranno et al., 2014; Zosel et al., 2020b) (hence smFRET can be used to validate results determined in isolation or detect the modulation of conformational preferences and/or structural dynamics through so-called quinary interactions [Guin and Gruebele, 2019]), and

  • high specificity for residues/domains via specific labeling. Biomolecules can be specifically labeled by a unique dye pair enabling smFRET measurements to be applicable on all sizes of molecules, including large complex assemblies (see Figure 1 [Kilic et al., 2018]), active biological machines (e.g., the ribosomes) (Dunkle et al., 2011) and even on whole native virions (Lu et al., 2019; Munro et al., 2014).

Workflow of modeling dynamic structures from FRET measurements.

(A) Integrative modeling requires structural and dynamic information. Prior information from conventional approaches (X-ray, NMR, cryoEM) together with computational tools defines the space of possible solutions for FRET-assisted structural modeling. The combination of structural (inter-dye distances) and dynamic information (kinetic connectivity and exchange rates) enables identification of a consistent model. (B) Study of structure and dynamics of chromatin fibers. A combined TIRF and confocal FRET study of structure and dynamics of chromatin fibers using three FRET labeling positions (DA1-3) for two pairs of dyes with distinct Förster distances. Förster distances ( is defined in section Inter-dye distances, Equation 6). Prior structural information provided by cryo-electron microscopy (top, left) (Song et al., 2014) and X-ray crystallography (top, right PDB ID: 1ZBB Schalch et al., 2005) is combined with the structural and dynamic information obtained by FRET experiments on immobilized molecules measured by total internal reflection (TIRF) microscopy and on freely diffusing molecules by confocal microscopy (Kilic et al., 2018). From the combined information, a consistent model is derived for chromatin fiber conformations with shifted registers, which are connected by slow (>100 ms) and fast de-compaction processes (150 µs) that do not proceed directly, but rather through an open fiber conformation. Figure 1B was reproduced from Figures 1, 3, and 6 in Kilic et al., 2018, Nature Communications with permission, published under the Creative Commons Attribution 4.0 International Public License (CC BY 4.0; https://creativecommons.org/licenses/by/4.0/).

© 2018, Kilic et al. Panel B was reproduced from Figures 1, 3 and 6 in Kilic et al., 2018 , with permission, published under the Creative Commons Attribution 4.0 International Public License.

Several methods have been utilized to determine structural ensembles such as NMR, single-particle cryoEM or XL-MS, and, recently, also smFRET in an integrative/hybrid (I/H) approach with computational modeling to overcome the sparsity of experimental data with respect to an atomistic description (Berman et al., 2019; de Souza and Picotti, 2020; Dimura et al., 2020; Gauto et al., 2019; Koukos and Bonvin, 2020; Na and Paek, 2020; Tang and Gong, 2020; Webb et al., 2018). I/H structural models derived from smFRET experiments using inter-dye distances as restraints were reported for flexible folded proteins (Brunger et al., 2011; Hellenkamp et al., 2017; Margittai et al., 2003; McCann et al., 2012), conformational ensembles of disordered/unstructured and unfolded proteins (Borgia et al., 2018; Holmstrom et al., 2018; Schuler et al., 2020), nucleic acids and protein-nucleic acid complexes (Craggs et al., 2019; Craggs and Kapanidis, 2012; Kalinin et al., 2012; Lerner et al., 2018b; Muschielok et al., 2008; Wozniak et al., 2008).

A further unique aspect of smFRET studies is that structural, kinetic, and spectroscopic information on large and complex systems can be recorded simultaneously in a single measurement. This facilitates linking dynamic and structural information in an integrative approach to (Figure 1A) (Hellenkamp et al., 2017; Kilic et al., 2018; Li et al., 2020b; Sanabria et al., 2020; Wasserman et al., 2016; Yanez Orozco et al., 2018):

  • define the number of possible structures consistent with data,

  • potentially reduce the ambiguity between different structural models compatible with the experimental data, and

  • reveal the dynamic exchange pathways that are structurally allowed.

As an example, Figure 1B shows the outcome of a multimodal smFRET study on the conformational landscape of a 12-mer chromatin array (~2.5 MDa) (Kilic et al., 2018) with dynamics occurring on timescales from nanoseconds to hours. SmFRET experiments could detect the flexible chromatin conformations (Figure 1B, middle panel), revealing their dynamic structural heterogeneity (Figure 1B, bottom panel), in contrast to the well-ordered static structures of chromatin fibers (Figure 1B, top panel). These flexible, partially-open and open conformations that are quite abundant in solution (population of >70%; Figure 1B, bottom panel) were not resolved before, although they are essential for proper gene organization and function. They represent the central interconversion hub for the distinct stacking registers of chromatin and are difficult to detect with other structural techniques. This approach of visualizing biomolecules in action under ambient conditions emphasizes the importance of their dynamic nature by resolving transitions between various conformational states, which, in many cases, promotes function (Aviram et al., 2018; Henzler-Wildman et al., 2007; Iljina et al., 2020; Lerner et al., 2018b; Sanabria et al., 2020; Tassis et al., 2020).

SmFRET measurements are typically performed using two approaches: with surface-immobilized molecules using total internal reflection fluorescence microscopy (TIRFM) and camera-based detection, or with freely diffusing molecules in solution using confocal microscopy and point detectors. Experimental systems are available commercially but are typically home-built. Samples are prepared and the data collected using lab-specific protocols, where data are stored in a variety of file formats and analyzed using an array of increasingly powerful software. For the field in general and for structural studies in particular, it is important to demonstrate that smFRET, as a method, is reproducible and reliable regardless of where and how the sample is measured. To this end, in an effort led by Thorsten Hugel, twenty laboratories joined in measuring smFRET on several dsDNA constructs (Hellenkamp et al., 2018a). Studying six distinct samples with different dyes and varying inter-dye distances, the mean FRET efficiencies obtained by the participating labs exhibited a surprisingly high degree of agreement (a ΔE between 0.02 and 0.05 depending on the details of the sample). The quantitative assessment and reproducibility of the intensity-based smFRET measurements and discussions about data analysis was an important milestone. These dsDNA FRET standards are now available for every day calibration and are especially useful for new groups joining the community.

Encouraged by the insights gained in the above-mentioned FRET endeavor (Hellenkamp et al., 2018a), new multi-lab blind studies have been initiated. The next comparative FRET study, led by Thorben Cordes, investigates the robustness and reliability of smFRET experiments on proteins undergoing ligand-induced conformational changes (Gebhardt et al., in preparation). This study uses two distinct model proteins to assess the reproducibility and accuracy of protein-based smFRET for inter-dye distance determination measurements. Protein systems bring new challenges, including statistical dye labeling, site-specific dye properties, protein stability, shipping, storage and conformational dynamics. Hence, the study also assesses the ability of smFRET to discover and quantify dynamics on different timescales from microseconds to seconds. Another FRET challenge, initiated by Sonja Schmid, is the kinSoftChallenge (http://www.kinsoftchallenge.com, Götz et al., in preparation), which evaluates existing tools for extracting kinetic information from single-molecule time trajectories. This challenge aims to: (1) demonstrate the ability of smFRET-based kinetic analyses to accurately infer dynamic information and (2) provide the community with the means of evaluating the different available software tools.

One important outcome of the various multi-lab FRET studies was that, although the agreement was good, it could be improved even further. In particular, the data analysis, and specifically corrections, can have an impact on the determined FRET efficiencies and resulting distances. Hence, an open discussion regarding which approaches work most reliably under what conditions is necessary. Access to the primary data and the ability to process them with various analysis approaches is, and will remain, the most transparent way to move the field forward. Currently, this is difficult given the many variations in methods employed, their documentation, file formats and experimental procedures implemented across laboratories establishing the optimal conditions, workflow and best practices even for existing, well-tested methods is challenging since a comparison of these methods is time-consuming and the necessary information is, in many cases, not available. With the increase in open scientific practices and submission of published data to repositories, a consensus is needed regarding what data and metadata should be stored and in which possible formats so that it can be readily utilized by the community.

Due to these considerations and the many opportunities for growth of the smFRET community, several laboratories with expertise in FRET, without pretension to be exhaustive or exclusive, have gathered to endorse these efforts and propose steps to organize the community around consistent and open-science practices. This action translates into general methodological recommendations or suggestions, which we introduce following the typical workflow of a smFRET experiment, including sample preparation and characterization, setup description, data acquisition and preservation, and data analysis. These recommendations on how to ‘practice’ smFRET are not an attempt to regiment the community but rather an initial suggestion that aims at encouraging an open dialog about existing practices in our field and leads to higher reproducibility in the results from smFRET experiments. We then discuss open science practices as well as the first steps that have been taken to form an international FRET community. We end with highlighting a few of the areas where we see smFRET making a big impact in various scientific fields in the near future.

State of the art of single-molecule FRET experiments

Within the FRET community, considerable know-how and expertise exists for the design, measurement and analysis of FRET experiments. In this section of the paper, we:

  • review the workflow of smFRET experiments,

  • discuss practical problems and potential pitfalls,

  • provide recommendations for good practice, and

  • list key scientific challenges that the field faces.

In the following, we consider each of these four aspects at every step of the smFRET workflow, from the choice of instrumentation all the way to the generation of structural and dynamic models.

Experimental approaches: free diffusion or surface immobilization?

The workflow of smFRET studies starts with choosing one of the two most popular smFRET implementations: confocal and TIRF microscopy. Confocal microscopy is especially well-suited for studying freely diffusing molecules (Figure 2A), while TIRF microscopy is typically used for surface-immobilized molecules (Figure 2B; e.g., reviewed in Juette et al., 2014; Roy et al., 2008; Sasmal et al., 2016).

Different smFRET modalities.

(A) Confocal smFRET measurements on freely-diffusing molecules. (i) A schematic of a single-color excitation confocal microscope with point detectors used for two-color detection. The excitation light is guided to the microscope body and reflected by a dichroic mirror (DM) toward a high numerical aperture (NA) objective lens that focuses the light in solution. The fluorescence emission is collected through the same objective lens, passes through the DM and pinhole and is spectrally split into donor and acceptor detection channels by a second DM in the detection path. After passing through emission filters (EF), single photons are detected on point detectors with high quantum efficiency, typically avalanche photodiodes (APD). (ii) Illustration of a double-labeled molecule freely diffusing through the confocal excitation spot. (iii) Exemplary confocal smFRET measurement showing photon bursts arising from single-molecules diffusing through the confocal volume. Green: Donor emission. Red: Acceptor emission. Exemplary bursts belonging to a single- or a double-labeled molecule are indicated with arrows. (iv) In ALEX or PIE experiments, the two-dimensional histogram of the molecule-wise FRET efficiency E and stoichiometry S allows one to separate single- and double-labeled populations (2005 Elsevier Ltd. All rights reserved. The figure was originally published as Figure 2A in Lee et al., 2005. Biophysical Journal, 88(4): 2939–2953. Further reproduction of this panel would need permission from the copyright holder). (B) TIRF-based smFRET experiments on surface-immobilized molecules. (i) Illustration of a surface-immobilized sample labeled with donor and acceptor fluorophores. (ii) Scheme of a single-color objective-type TIRF excitation two-color wide-field detection microscope. A: Aperture, TL: Tube lens, L: Lens, M: Mirror, DM: Dichroic mirror, EF: Emission filter. (iii) Illustration of an image of single molecules, in which the donor and acceptor (FRET) signals are split onto two halves of the camera. Mapping between the two channels is typically done using fluorescent beads (Joo and Ha, 2012; Roy et al., 2008; Zhuang et al., 2000) or zero-mode waveguides (Salem et al., 2019). (iv) Single-molecule fluorescence trajectory of the donor and acceptor (FRET) dyes, illustrating an anti-correlation indicative of FRET dynamics.

© 2005, Elsevier. All rights reserved. Panel Aiv was originally published as Figure 2A in Lee et al., 2005. Further reproduction of this panel would need permission from the copyright holder.

Compared to most other single-molecule approaches, both smFRET modalities offer relatively high throughput.

  • In the confocal modality, the free diffusion of molecules into the observation volume and the short residence times enable the acquisition of many single-molecule events for extended amounts of time at rates of a few events per second. It can offer sub-nanosecond time resolution, yet single molecules are only observed during diffusion through the confocal excitation volume (typically <10 milliseconds). This allows one to obtain snapshots of thousands of individual molecules over the course of hours.

  • In the TIRF modality, hundreds to thousands of dye-labeled molecules can be imaged simultaneously in one field of view. This approach reveals ‘motion pictures’ of individual molecules from seconds to minutes until the fluorophores photobleach. It typically has a lower temporal resolution of about a few tens of milliseconds but this is improving with technological advances. TIRF can be performed by illuminating through a high-numerical-aperture objective (Figure 2B) or through a quartz prism (Roy et al., 2008).

When embarking on the investigation of conformational dynamics of a new biological system, the method of choice most often depends on the availability of the proper instrumentation. However, the dynamical aspects (reviewed in section Conformational dynamics) of the biological system under investigation, which are typically not known a priori, will eventually define which of the two methods is best suited. Because the dynamics of biological systems occur over a range of timescales from nanoseconds to seconds (Figure 3), ideally one would like to apply both modalities in parallel to obtain a complete understanding of the system (e.g., as shown in Figure 1).

Exemplary methods for following smFRET dynamics on different timescales.

Top: Biomolecular dynamics cover a wide range of timescales. Biomolecular rotations occur in the pico- to nanosecond range, while conformational changes take place in nano- to microseconds (ns-µs), as in chain dynamics of disordered proteins, and protein folding in microseconds to minutes. Transitions along energetically unfavorable pathways can take up to hours or longer, as in protein misfolding (Borgia et al., 2011; Tosatto et al., 2015). (2013 Elsevier Ltd. All rights reserved. The figure was originally published as Figure 1 in Schuler and Hofmann, 2013. Current Opinion in Structural Biology, 23(1): 36–47. Further reproduction of this panel would need permission from the copyright holder.) Bottom: (A) Picosecond (ps) to millisecond (ms) processes are typically examined with confocal methods such as polarization-resolved fluorescence lifetime measurements and Fluorescence Correlation Spectroscopy (FCS). Example shown: chain dynamics of an IDP from nsFCS. (B) Conformational states are identified by individual populations with characteristic positions in the FRET efficiency - lifetime diagrams as discussed in the sections Detection and characterization of intra-state dynamics and Future of smFRET (adapted from Soranno et al., 2012). (C) Fast transitions measured using confocal microscopy can be analyzed using the photon trajectory and applying a photon-by-photon maximum likelihood approach (2018 Elsevier Ltd. All rights reserved. The figure was originally published as Figures 2 and 3 in Chung and Eaton, 2018. Current Opinion in Structural Biology, 48: 30–39. Further adaptation of this panel would need permission from the copyright holder.) The timescale over which kinetics can be measured can be extended for diffusing molecules at low concentrations by using a recurrence analysis of single particles (RASP, Hoffmann et al., 2011). (D) Non-equilibrium experiments over extended periods of time can be performed with microfluidic mixing devices. (Copyright 2011, Nature Publishing Group, a division of Macmillan Publishers Limited. All Rights Reserved. Reproduced from Gambin et al., 2011, with permission. Nature Methods 8:239–241. Further reproduction of this panel would need permission from the copyright holder.) (E) Slow changes in conformations over a broad range of timescales can be followed in smFRET efficiency trajectories registered by single-photon counting (SPC) or cameras over minutes to many hours when the sample is immobilized (adapted from Figure 1 of Zosel et al., 2018).

© 2013, Elsevier Ltd. All rights reserved. Figure 3 (top) and panel A was originally published as Figure 1 in Schuler and Hofmann, 2013. Further reproduction of this panel would need permission from the copyright holder.

© 2018, Elsevier Ltd. All rights reserved. Panel C was originally published as Figures 2 and 3 in Chung and Eaton, 2018. Further adaptation of this panel would need permission from the copyright holder.

© 2011, Nature Publishing Group, a division of Macmillan Publishers Limited. All Rights Reserved. Panel D was originally published as Figure 1f in Gambin et al., 2011. Further reproduction of this panel would need permission from the copyright holder.

Many variations exist with respect to the above-mentioned basic modalities to:

1) maximize the information content of the fluorescence signal.

2) optimize data collection.

3) control the sample.

The many possibilities available in the choice of hardware underscore the importance of precisely describing the components of the experimental setup. This includes optical elements (e.g., lenses, filters, mirrors, dichroics), light sources, optomechanical/optoelectronic devices and their characteristics, and detectors and their associated electronics. These details contribute in many ways to the finally recorded data and cannot, in general, be inferred retrospectively.

With the palette of FRET modalities increasing steadily, we recommend a rigorous comparative study of the different methods using well-characterized model samples. First and foremost, the study should determine the precision and limitations of each method and their complementarity. As one example, potential pitfalls in the determination of data correction factors (described in the section FRET efficiency) could be identified by a side-by-side comparison of fluorescence lifetime and intensity-based FRET methods.

Sample preparation

Dyes

For studying biomolecular conformations and their dynamics with smFRET, the biomolecules of interest must be labeled with organic dyes that are suitable for single-molecule fluorescence detection (intrinsically fluorescent aromatic amino acids are not stable or bright enough). These dyes usually include three modules: (i) a chemically reactive group that forms a covalent bond preferentially with a specific nucleic acid base or amino acid residue of choice, (ii) a sufficiently long linker of a few connecting bonds to ensure isotropic rotation of the fluorophore, and (iii) an (often bulky) π-conjugated fluorophore that typically has hydrophobic regions and charged or polar substitutions.

To compete with background-noise, smFRET-compatible dyes should be very bright. They should hence possess a sufficiently large extinction coefficient (>50,000 M−1cm−1 at the wavelength of excitation) and high fluorescence quantum yield (ϕF ≳ 0.3), be very photostable (≳ 106 excitation cycles before photobleaching), exhibit low photoblinking, should not possess long-lived dark states to avoid optical saturation and have a large fundamental anisotropy, that is have approximately collinear absorption and emission transition dipole moments (typically, r0≳0.37). The fluorescence lifetime should be on the 1-5 ns scale. In the case of TCSPC experiments, a general rule of thumb is that the laser repetition period should be chosen at least four times as large as the fluorescence lifetime. For instance, for a dye with a fluorescence lifetime of 4 ns, a laser pulse repetition rate of ~64 MHz for one-color excitation or ~32 MHz for two-color nsALEX/PIE experiments should be used. In addition, using dyes with intrinsic mono-exponential fluorescence decays simplifies the analysis. Continuous efforts are ongoing to further improve smFRET dyes by:

Finally, a pair of FRET dyes should always be chosen such that its Förster distance, R0, (defined in section Inter-dye distances, Equation 6) is around the expected inter-probe distance, RDA, where the dependence of the FRET efficiency, E, is most sensitive to RDA. When quantifying conformational dynamics, the FRET dye pair should be chosen such that the expected change in FRET efficiency is as large as possible.

Conjugation

To measure intra-molecular distances within biomolecules, smFRET experiments require the conjugation of two dye molecules to the same biomolecule or the same biomolecular complex. Site-specific conjugations in proteins utilize the introduction of point mutations, typically to cysteines, that will accommodate the specific conjugation chemistry, usually maleimide- or iodoacetamide-cysteine chemistry. In this case, two cysteines are often stochastically labeled, leading to a mixture of donor-acceptor and acceptor-donor labeled molecules. While interchanging the donor and acceptor positions has a negligible effect, from the geometric standpoint, on the FRET-averaged distance (Peulen et al., 2017), stochastic labeling might cause problems when the donor/acceptor dyes possess different spectroscopic properties at the different labeling positions.

Potential issues related to stochastic labeling can be excluded when, for example, a multi-dimensional analysis available from MFD-PIE shows no dye-induced sub-populations. Alternatively, stochastic labeling can also be avoided by:

Different approaches are applied for nucleic acids (e.g., reviewed in Hanspach et al., 2019; Steffen et al., 2019). For short nucleic acids, site-specific conjugation is generally achieved by post-synthetic labeling of reactive groups (e.g., through click chemistry) that are incorporated during solid-phase synthesis. Strategies have also been developed to site-specifically label longer RNAs (Anhäuser and Rentmeister, 2017; Baum and Silverman, 2007; Büttner et al., 2014; Zhao et al., 2018), and the use of hybridizing probes (Steiner et al., 2008) and fluorescent nucleobase analogues as intrinsic probes (Karimi et al., 2020; Steinmetzger et al., 2020) has been explored.

A general recommendation for labeling is to aim for high-purity sample preparations with optimized labeling protocols, as only this will result in substantially and specifically labeled samples with both donor and acceptor dyes. Single-molecule measurements have the ability to separate out the donor-acceptor-labeled molecules and thus purify the sample ex post facto, but a significant amount of double-labeled samples is advantageous. After labeling, we recommend using a rigorous screening procedure that compares the activities of labeled and unlabeled wild-type biomolecules to determine whether the mutations introduced to a biomolecule and/or the labeling with the dyes significantly influence the biomolecule's functionality (e.g., catalytic activity, binding affinity) and stability (e.g., against denaturants or thermally-induced transition curves) (Best et al., 2018; Deniz et al., 2000; Lerner et al., 2018b; Orevi et al., 2014; Riback et al., 2019; Sottini et al., 2020). To check for structural integrity, methods such as mass spectrometry, circular dichroism (CD), dynamic light scattering (DLS), and small-angle X-ray scattering (SAXS) can be used (Best et al., 2018; Borgia et al., 2016; Riback et al., 2019). We also recommend reporting the labeling and purification procedures as well as the labeling efficiency. In cases where no labeling alternative exists that does not modify the structure and/or rate of function, mechanistic insights into biomolecules or complexes can often still be obtained. Nevertheless, the results and conclusions concerning wild-type and unlabeled protein, respectively, should be interpreted cautiously. Finally, when samples need to be frozen/thawed, we recommend testing the long-term stability and functionality versus fresh protein preparations.

Immobilization

For long observation times, labeled molecules are typically immobilized. This is most frequently achieved via a biotin-streptavidin linkage. Immobilization must be carefully performed in order to systematically eliminate spurious contributions from molecules that are non-specifically bound (Lamichhane et al., 2010; Traeger and Schwartz, 2017). To address this potential issue, efforts have been made to optimize surface passivation procedures (Hua et al., 2014; Kuzmenkina et al., 2005; Park et al., 2020; Selvin and Ha, 2008). Alternatives that avoid the direct linking of biomolecules to surfaces are:

We recommend reporting the immobilization conditions, the control experiments that demonstrate the specific nature of the surface immobilization strategy, and the percentage of functional or dynamic molecules (Bavishi and Hatzakis, 2014; Lamichhane et al., 2010; Roy et al., 2008) in detail. Finally, when possible, we recommend cross-validating the results of surface-immobilization based smFRET experiments by comparing them either to those obtained in ensemble or single-molecule FRET experiments on non-immobilized, freely-diffusing molecules (Pirchi et al., 2011), or to results using different immobilization strategies (Gregorio et al., 2017; Whitford et al., 2010).

Spectroscopic characterization

Fluorescent dyes are characterized by particular spectroscopic properties, which may change when conjugated to a protein (Lerner et al., 2013; Peulen et al., 2017; Sindbert et al., 2011; Steffen et al., 2016) or even between different structural states of the labeled biomolecule (Kudryavtsev et al., 2012). The most important artifacts to look out for are:

When the local and/or global environment influences the photophysical properties of either the donor or the acceptor dyes differently, different subpopulations might appear (Kalinin et al., 2010a; Vandenberk et al., 2018). Depending on the research question at hand, these subpopulations per se may provide additional information beyond FRET (e.g., PIFE [Ploetz et al., 2016], PET [Doose et al., 2009], or quenchable FRET [Cordes et al., 2010]). In cases where accurate distance measurements are needed, properly designed control experiments of fluorescence lifetimes and anisotropies of single-label versions for both labeling positions and dyes can be used to detect and eventually correct these spectroscopic alterations a posteriori. In addition, dye-artifacts can be identified from the information provided by ALEX or PIE experiments (Kapanidis et al., 2004; Kudryavtsev et al., 2012), MFD-based detection (Hellenkamp et al., 2017; Rothwell et al., 2003) or analysis of the width of FRET efficiency distributions (Kalinin et al., 2010a; Nir et al., 2006). Note that the influence of dye photoblinking must be taken into account: (1) when determining the correction factors necessary for precise FRET efficiency measurements (see section Determining absolute FRET efficiencies from fluorescence intensities) or (2) in the donor fluorescence quantum yield, when accurate distance estimations are required, which, in turn, depends on a correct Förster distance, R0 (defined in section Inter-dye distances, Equation 6).

When dye- and microenvironment- dependent influences exist, they can be characterized or taken into account by a careful choice of fluorophores and/or labeling locations or coarse-grained computer simulations (Peulen et al., 2017), or they can be ruled out completely by validating the observations with (an)other FRET pair(s) (Borgia et al., 2018; Borgia et al., 2016; de Boer et al., 2019b; Husada et al., 2018; Lerner et al., 2017; Vandenberk et al., 2018; Voelz et al., 2012) or switching fluorophore positions (Sanabria et al., 2020). How important a detailed spectroscopic analysis is, depends on the nature of the research question being addressed.

Photostabilization

Often, chemical photostabilizers are added to reduce oxidative photodamage by lowering the time spent in triplet or radical-ion dark states (Ha and Tinnefeld, 2012; Widengren et al., 2007). The choice of the photostabilizing agent is specific to the fluorophore used and finding the correct conditions for both the donor and acceptor fluorophores can be challenging. Commonly used photostabilizers for smFRET include 6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid (Trolox) (Cordes et al., 2009; Dave et al., 2009; Rasnik et al., 2006; Vandenberk et al., 2018), n-propylgallate (Widengren et al., 2007), β-mercaptoethanol (Campos et al., 2011; Ha and Tinnefeld, 2012), ascorbic acid (Aitken et al., 2008; Gidi et al., 2020; Vogelsang et al., 2008; Widengren et al., 2007), linear polyenes (Pfiffi et al., 2010) and cyclopolyenes (Dave et al., 2009; Targowski et al., 1987; Widengren et al., 2007), methylviologen (Vogelsang et al., 2008) and a range of other compounds (Glembockyte et al., 2015; Isselstein et al., 2020). For optimal performance, reducing and oxidizing agents can be combined (Dave et al., 2009; Vogelsang et al., 2008). Fluorophore performance and photon budgets can be enhanced by removing oxygen from the buffer through oxygen scavenging systems such as glucose oxidase (Kim et al., 2002) or the PCA/PCD system (Aitken et al., 2008), in which case an exogenous triplet quencher, such as those mentioned above, is required to prevent long-lived dark states. In any case, we recommend verifying that the use of these photostabilization reagents does not interfere with the biological system under study. In the case of lipid bilayers, an influence of several of the commonly used photostabilization agents on membrane properties was observed (Alejo et al., 2013).

Molecule identification and validation

After data collection in either confocal or TIRF modalities, the single-molecule fluorescent signal in the resulting time traces or videos must be identified and validated before further detailed analysis can be performed.

Identification

In the confocal modality, the raw ‘burst’ data includes a sequence of photon detection or arrival times from at least two detectors. The first step is to identify fluorescence bursts arising from single molecules from the background, commonly referred to as the ‘burst search’ (Figure 2A–iii). Various approaches have been described for the robust and accurate detection of single-molecule events (Enderlein et al., 1997; Fries et al., 1998; Nir et al., 2006; Schaffer et al., 1999; Sisamakis et al., 2010). After the burst search step, the identified single-molecule events are filtered based on the burst properties (e.g., burst size, duration or width, brightness, burst separation times, average fluorescence lifetime or quantities calculated from these burst parameters). The burst search and burst selection criteria have an impact on the resulting smFRET histograms. Hence, we recommend that the applied burst property thresholds and algorithms should be reported in detail when publishing the results, for example, in the methods section of papers but potentially also in analysis code repositories. Often, burst search parameters are chosen arbitrarily based on rules-of-thumb, standard lab practices or personal experience. However, the optimal burst search and parameters vary based on the experimental setup, dye choice and biomolecule of interest. For example, the detection threshold and applied sliding (smoothing) windows should be adapted based on the brightness of the fluorophores, the magnitude of the non-fluorescence background and diffusion time. We recommend establishing procedures to determine the optimal burst search and filtering/selection parameters.

In the TIRF modality, molecule identification and data extraction can be performed using various protocols (Börner et al., 2016; Holden et al., 2010; Juette et al., 2016; Preus et al., 2016). In brief, the molecules first need to be localized (often using spatial and temporal filtering to improve molecule identification) and then the fluorescence intensities of the donor and acceptor molecules extracted from the movie. The local background needs to be determined and then subtracted from the fluorescence intensities. Mapping is performed to identify the same molecule in the donor and acceptor detection channels. This procedure uses a reference measurement of fluorescent beads or zero-mode waveguides (Salem et al., 2019) or is done directly on samples where single molecules are spatially well separated. The outcome is a time series of donor and acceptor fluorescence intensities stored in a file that can be further visualized and processed using custom scripts. In a next step, filtering is generally performed to select molecules that exhibit only a single-step photobleaching event, that have an acceptor signal when the acceptor fluorophores are directly excited by a second laser, or that meet certain signal-to-noise ratio values. However, potential bias induced by such selection should be considered.

User bias

Despite the ability to manually determine burst search and selection criteria, molecule sorting algorithms in the confocal modality, such as those based on ALEX/PIE (Kapanidis et al., 2005; Kudryavtsev et al., 2012; Tomov et al., 2012), do not suffer from a substantial user bias. In the early days, many TIRF modality users have relied on visual inspection of individual single-molecule traces. Such user bias was considerably reduced by the use of hard selection criteria, such as intensity-based thresholds and single-step photobleaching, intensity-based automatic sorting algorithms (e.g., as implemented in the programs MASH-FRET [Hadzic et al., 2019], iSMS [Preus et al., 2015] or SPARTAN [Juette et al., 2016]), and, most recently, artificial intelligence-based molecular sorting (deepFRET [Thomsen et al., 2020] and AutoSiM [Li et al., 2020a]).

Single-molecule experiments are often advertised as being able to detect rare events. Nonetheless, even for such sparsely populated states, it has to be confirmed that they are biologically relevant and neither a result of the selection procedure, coincidence or photophysical artifacts. To this end, users should specify how selections were performed and what percentage of the molecules was used for further analysis.

Ideally, a recommended protocol with implicit validation would be to start in the confocal modality to determine (i) the degree of labeling, (ii) the FRET properties of major biochemical species, and (iii) their populations and dynamic properties (see Figure 1). With this information at hand, experiments can be performed in the TIRF modality, where the percentage of FRET-active molecules and their FRET properties can be directly compared with the confocal data. Both datasets should be mutually consistent and, in this way, provide direct feedback with respect to potential artifacts (e.g., due to immobilization).

Conformational dynamics

Many users in the FRET community employ the detection and characterization of different subpopulations or measurements of conformational dynamics as a handle to study biomolecules or biomolecular systems. Conformational dynamics are typically defined as:

  • conformational transitions between distinct states separated by an activation barrier, typically defined as larger than the thermal energy, kBT, where kB is Boltzmann’s constant and T is the absolute temperature, and

  • or conformational fluctuations within states, defined by the shape of the potential wells between activation barriers.

Transitions can occur under equilibrium conditions, can be induced by the addition of substrates, ligands, or interaction partners (de Boer et al., 2019a; Mapa et al., 2010; Mazal et al., 2018; Schluesche et al., 2007); induced by mixing with denaturants (Kuzmenkina et al., 2006; Lindhoud et al., 2015; Maity and Reddy, 2016; Moosa et al., 2018; Nienhaus, 2006; Pirchi et al., 2011; Rieger et al., 2011; Schuler et al., 2002); or triggered by temperature (Ebbinghaus et al., 2010; Holmstrom et al., 2014; Nettels et al., 2009; Zhao et al., 2010a) and pressure modulations (Schneider et al., 2018; Sung and Nesbitt, 2020). Structural transitions can also occur spontaneously.

SmFRET is unique in that it allows the detection and analysis of equilibrium and non-equilibrium conformational dynamics across at least 12 orders of magnitude in time, that is from the nanoseconds to, in principle, thousands of seconds (Figure 3). Notably, it is important to optimize the labeling positions to maximize the distinction between different conformational states based on their FRET efficiencies (Dimura et al., 2020).

Detecting dynamics

Biomolecules are dynamic systems that show conformational flexibility and dynamics on fast time scales (Henzler-Wildman and Kern, 2007). Oftentimes, conformational interconversions occur on a timescale faster than the sampling time of the detection system, for example < 10 ms for TIRF modality or < 0.1 ms for confocal modality, resulting in the observed single-molecule time series or FRET efficiency histogram exhibiting only time-averaged FRET values, weighted by the fractional population of each conformational state. Several groups have developed methods for detecting and analyzing such ‘dynamic averaging’ from confocal-modality data. In general, these methods allow retrieval of dynamics on the milliseconds and sub-millisecond timescales by analyzing the average fluorescence lifetimes and/or photon counting statistics of single-molecule bursts. The precise knowledge of the experimental shot noise separates smFRET from other techniques in structural biology and enables a quantitative analysis of fluctuations caused by biomolecular dynamics. A number of methods have been developed for detecting and quantifying smFRET dynamics, which we discuss in more detail below on slower (section Slow dynamics) and faster time scales (section Faster dynamics). The first step in analyzing smFRET dynamics is the verification that dynamics are present. Popular methods for the visual detection of dynamics include:

Slow dynamics

For dynamics on the order of 10 ms or slower, transitions between conformational states can be directly observed using TIRF-modality approaches, as have been demonstrated in numerous studies (Blanchard et al., 2004; Deniz, 2016; Juette et al., 2014; Robb et al., 2019; Sasmal et al., 2016; Zhuang et al., 2000). Nowadays, hidden Markov models (HMM) (Figure 4E) are routinely used for a quantitative analysis of smFRET time traces to determine the number of states, the connectivity between them and the individual transition rates (Andrec et al., 2003; Keller et al., 2014; McKinney et al., 2006; Munro et al., 2007; Steffen et al., 2020; Stella et al., 2018; Zarrabi et al., 2018). Below, we list extensions and other approaches for studying slow dynamics.

Exemplary selection of approaches to detect and quantify conformational dynamics in smFRET.

(A) Dynamics in a three-state system are detected using the two-dimensional distribution of the FRET efficiency and donor fluorescence lifetime (Reproduced from Gopich and Szabo, 2012. Further reproduction of this panel would need permission from the copyright holder.) (B) The two-state dynamics of a DNA hairpin are revealed from the standard deviation of the proximity ratio E that is higher than expected for photon counting statistics alone in the burst variance analysis (BVA). (2011 The Biophysical Society. Published by Elsevier Inc All rights reserved. The figure was originally published as Figure 4C in Torella et al., 2011. Biophysical Journal, 100(6): 1568–1577. Further reproduction of this panel would need permission from the copyright holder.) (C) In fluorescence correlation spectroscopy (FCS), the dynamics show up as a positive correlation in the autocorrelation functions GDD and GAA and an anti-correlation in the cross-correlation function GDA (2010 American Chemical Society Ltd. All rights reserved. The figure was originally published as Figure 1B and C in (Gurunathan and Levitus, 2010, reproduced with permission). Copyright 2010 ACS Publications. Further reproduction of this panel would need permission from the copyright holder.) (D) Photon-by-photon maximum likelihood estimation (MLE) infers the kinetic parameters directly from the photon arrival times (2011 American Chemical Society Ltd. All rights reserved. The figure was originally published as the Abstract Figure in Chung et al., 2011, reproduced with permission. Copyright 2011 ACS Publications. Further reproduction of this panel would need permission from the copyright holder.) (E–F) A hidden Markov model (HMM) is applied to the time traces of the FRET efficiency to estimate the states and interconversion rates (E). From the transition density plot (F), the connectivity of the FRET states is revealed. Displayed data in E and F are simulated. (Panels E and F: 2006 The Biophysical Society. Published by Elsevier Inc All rights reserved. The figures were originally published as Figure 4A and D in McKinney et al., 2006. Biophysical Journal, 91(5): 1941–1951. Further reproduction of this panel would need permission from the copyright holder.)

© 2012, Gopich and Szabo. Panel A was originally published as Figure 1C in Gopich and Szabo, 2012. Further reproduction of this panel would need permission from the copyright holder.

© 2011, The Biophysical Society. Published by Elsevier Inc. All rights reserved. Panel B was originally published as Figure 4C in Torella et al., 2011. Further reproduction of this panel would need permission from the copyright holder.

© 2010, American Chemical Society Ltd. All rights reserved. Panel C was originally published as Figures 1B and 1C in Gurunathan and Levitus, 2010. Further reproduction of this panel would need permission from the copyright holder.

© 2011, American Chemical Society Ltd. All rights reserved. Panel D was originally published as the abstract figure in Chung et al., 2011. Further reproduction of this panel would need permission from the copyright holder.

© 2006, The Biophysical Society. Published by Elsevier Inc. All rights reserved. Panels E and F were originally published as Figures 4A and 4D in McKinney et al., 2006. Further reproduction of this panel would need permission from the copyright holder.

There are still many challenges with respect to the accuracy of the approaches that need to be discussed and improvements made to provide a reliable determination of kinetics.

Faster dynamics

Several methods exist that assist in the quantification of the kinetic parameters governing fast conformational dynamics, as also exemplified in Figure 3A,B.

Fluorescence correlation spectroscopy (FCS) (Magde et al., 1972; Rigler et al., 1993) methods have also been widely applied and the observed kinetic rates are model-independent (i.e., unbiased).

In the analysis of fast dynamics and subpopulations, it is generally important to account for fast photophysical transitions, such as dye photoblinking, and for interactions of the dyes with the biomolecular surface, which may interfere with subpopulation dynamics and result in inaccurate transition rates (Chung and Gopich, 2014; Ingargiola et al., 2018b; Lerner et al., 2018b).

The detection and analysis of fast dynamics is one of the issues addressed in the protein FRET challenge (Gebhardt et al., in preparation). Finally, at the moment of this writing, different analysis algorithms are typically being applied to the data independently from each other (e.g., MFD, BVA, 2CDE, PDA, filtered-FCS in burst analysis) while, in fact, they could corroborate each other or even help in deciding on models. The field could focus on creating global multi-algorithm workflows or tools to test, how a model obtained with one algorithm would influence the results of other analyses.

Detection and characterization of intra-state dynamics

Rapid structural dynamics within a given conformational state, that is within a single energy minimum, can also be studied with smFRET by modeling them as a continuous distribution rather than a state-dependent distribution. In the given example, rigid and flexible conformational states can be distinguished in the measurement. Information regarding the flexibility of a given conformation can be retrieved by:

It is, however, important to mention that the distinction between dynamics within a conformational state and dynamics of transitions between different conformational states is still under debate, which highly depends on the definition of an activation barrier for different modes of structural dynamics and on the different smFRET modalities used.

FRET efficiency

The efficiency of the energy transfer process, that is the FRET efficiency, E, is defined as the fraction of donor excitations that result in energy transfer. Assuming a single distance between the centers of the donor and acceptor molecules, RDA, the FRET efficiency is given by:

(1) E=kFRETkFRET+kD=11+(RDAR0)6,

where kFRET is the rate of energy transfer, kD is the rate of donor de-excitation in the absence of an acceptor molecule, and R0 is the Förster distance (discussed in section Dye models). Hence, FRET is indeed a tool that can measure distances on the molecular scale (Förster, 1948; Stryer and Haugland, 1967). For many smFRET studies, a qualitative indicator of the inter-probe distance is sufficient, for example, to merely be able to distinguish between conformational subpopulations or their transitions. Therefore, for all FRET experiments that do not require the exact inter-dye distance, the absolute value of E does not need to be known. However, special care should be taken to ensure that the observed changes of the donor and acceptor intensities report on a structural change of the molecule and are not a result of dye photophysics or dye-surface interactions. In the cases where accurate distance measurements are desired, smFRET can be used for that purpose.

Determining absolute FRET efficiencies from fluorescence intensities

Typically, in smFRET, the FRET efficiency is determined from the fluorescence intensities:

(2) E=FAem|DexFDem|Dex+FAem|Dex,

where FAem|Dex is the sensitized fluorescence signal from the acceptor after donor excitation and FDem|Dex is the signal emanating from the donor. Here, we use a notation specific to experiments using alternating laser excitation, but equivalent expressions can be derived for single-color excitation. In reality, the absolute value for E requires knowledge of some correction factors (Hellenkamp et al., 2018a; Lee et al., 2005):

(3) E=IAem|Dex -αIDem|Dex -δIAem|AexγIDem|Dex +IAem|Dex -αIDem|Dex -δIAem|Aex ,

where IAem|Dex is the background-corrected signal in the acceptor emission channel after donor excitation, IDem|Dex is the background-corrected signal in the donor emission channel after donor excitation and IAem|Aex is the background-corrected signal in the acceptor emission channel after acceptor excitation. The last term can be estimated using the acceptor-only species and fluorescence signal after acceptor excitation when the ALEX/PIE method is used (Hellenkamp et al., 2018a; Kudryavtsev et al., 2012; Lee et al., 2005) or by comparing fluorescence intensities before and after donor photobleaching prior to acceptor photobleaching in trajectories from immobilized molecules (Yoo et al., 2018).

The required correction factors are:

  • α, the fraction of the donor fluorescence signal detected in the acceptor channel due to spectral crosstalk,

  • δ, the fraction of acceptor photons arising from excitation of the acceptor at the wavelength of the donor-exciting laser, directly, and not excitation via energy transfer,

  • the γ factor (Ha et al., 1999), which compensates for the fact that the number of photons detected from the donor and acceptor fluorophores is not proportional to the number of their excitation/de-excitation cycles for two reasons: (i) fluorophores, in general, have different fluorescence quantum yields, ϕF values, and (ii) the efficiencies of detecting photons are different for the two channels due to different optical transmission efficiencies (owing to the characteristics of the filters and optics used) and different spectral sensitivities of the detectors.

The optimal procedures for determining the correction parameters is still a matter of active debate within the community. In the following, we focus on the γ factor, which we identify as the major contribution to uncertainty in smFRET experiments.

Determining the γ factor in confocal mode

Whenever a broad E distribution is reported in the confocal mode, the γ factor can be extracted using ALEX/PIE measurements. This method exploits the fact that the stoichiometry parameter, S (Kapanidis et al., 2004) (i.e., the ratio between the number of photons emitted after donor excitation and the number of photons emitted after donor and acceptor excitations), is independent of E when properly corrected for γ. It is thus essential that the sample contains two or more species with different distances and thus FRET efficiencies, E, yet identical values of γ for this method to work (Lee et al., 2005). Thus, accurate measurements of ϕF for both dyes have to be performed for each species. Alternatively, fluorescence lifetime measurements and the correlated analysis of intensity and lifetime data is often used to determine individual γ factors for each E sub-population, since lifetime-based FRET, in principle, provides the absolute E of a sub-population of single-molecule bursts independently from its intensity-based counterpart (Rothwell et al., 2003; Sisamakis et al., 2010; Vandenberk et al., 2018). However, when one or more species are dynamically averaged, a proper determination of the γ factor becomes more challenging and different assumptions need to be made.

Currently, the uncertainty in the determination of γ is one of the largest contributions to discrepancies of smFRET histograms measured from different laboratories (Gebhardt et al., in preparation). Hence, it would be beneficial to discuss optimal approaches to determine a robust confocal-mode γ value.

Determining the γ factor in TIRF mode

When ALEX data are collected on immobilized samples, the γ factor can also be estimated for individual molecules, provided that the acceptor photobleaches before the donor (Ha et al., 1999; Hildebrandt et al., 2015; McCann et al., 2010). Here, the decrease in the acceptor signal and the increase in donor signal upon acceptor photobleaching can be directly compared. This is also true for molecules undergoing slow dynamics between different conformations as the changes in intensity reflect the changes in detection efficiency. For this approach to be accurate, however, the acceptor must not enter a transient (e.g., redox or triplet) state that still absorbs energy from the donor (Hofkens et al., 2003; Nettels et al., 2015). The individual γ factors are usually broadly distributed, indicating a potential variability in its value. Nevertheless, an average γ factor is often applied to molecules where the donor photobleaches before the acceptor.

Determining absolute FRET efficiencies from fluorescence lifetimes

In addition to the traditional intensity-based FRET efficiency (E) can also be determined from the fluorescence lifetime (τD(A)) of the donor in the presence and absence of the acceptor, denoted by τD(A) and τD(0), respectively. Assuming a single distance between donor and acceptor fluorophores (i.e., no distance fluctuations), the FRET efficiency is given by:

(4) E=1τD(A)τD(0)

The advantage of this approach is that correction factors are not needed, as most of the above-mentioned corrections influence the relative number of photons detected in the donor and acceptor channels, but not the donor fluorescence decay. The lifetime approach can also be used in ensemble/imaging measurements under conditions of incomplete labeling. Combined intensity- and lifetime-based FRET efficiencies can additionally be used for checking the self-consistency of the data and for detecting dynamics (e.g., via E-τ plots) (Gopich and Szabo, 2012; Kalinin et al., 2010b; Rothwell et al., 2003; Schuler et al., 2016).

Other methods for determining FRET efficiencies

There are additional procedures for determining the FRET efficiency, most of which are compatible with single-molecule fluorescence techniques. The FRET efficiency can also be determined:

  • from the steady-state donor anisotropy (Clegg, 1992),

  • from the ratio of the acceptor’s intensity after donor excitation to the acceptor’s intensity after acceptor excitation (Clegg, 1992),

  • from the acceptor’s intensity in the presence and absence of the donor (e.g., via donor photobleaching) (Clegg et al., 1992),

  • from the donor’s intensity in the presence and absence of the acceptor (e.g., via acceptor photobleaching) (Bastiaens et al., 1996),

  • from time-resolved anisotropy measurements, in particular in homo-FRET experiments, where two identical probes are used as a donor-acceptor pair (Bergström et al., 1999; Somssich et al., 2015),

  • using fluorescence correlation spectroscopy methods (Müller et al., 2005; Widengren et al., 2001).

Inter-dye distances

When smFRET experiments are used for structural studies or accurate distance determination is desired, many steps need to be taken to convert the raw data (photons detected and registered by the detectors) into absolute inter-dye distances. In essence, it requires exact knowledge of the Förster distance, R0 (also referred to as the Förster radius) and therefore of all parameters required for determining it, as well as knowledge with respect to the flexibility of the attached fluorophores (approximated using a dye-model). In this section, we review the various issues involved.

Förster distance R0

In FRET, the excitation energy of the donor fluorophore is transferred to an acceptor fluorophore via weak dipolar coupling. Considering a single donor-acceptor distance, RDA, the efficiency, E, of this non-radiative transfer process scales with the sixth power of RDA normalized by the Förster distance, R0 (Equation 1). In smFRET studies, dyes are usually coupled to the biomolecules via long (ranging typically between 10 and 15 atoms) mostly flexible linkers, which result in an equilibrium distribution of RDA values, p(RDA), caused by the flexibility of the dye linkers. In this case, one may observe a mean FRET efficiency ⟨E⟩ related to the FRET efficiency, averaged over all distances and their probabilities:

(5) E=0p(RDA)1+(RDAR0)6dRDA.

It is noteworthy to mention that Equation 5 holds under the assumption that the inter-dye distance remains unchanged during the excited-state lifetime of the donor fluorophore. From the mean FRET efficiency ⟨E⟩, one obtains the FRET-averaged apparent donor-acceptor distance, RDAE, which differs from the distance between the mean dye positions (Kalinin et al., 2012) and is dependent on the flexibility and dynamics of the dye.

As mentioned before, R0 (Equation 1) is the distance at which half of the donor de-excitation events occur via energy transfer to the acceptor fluorophore. R0 (in Å) is given by:

(6) R0=0.2108(κ2ΦF,D(0)nim4FD(λ)εA(λ)λ4dλ)16,

meaning that it depends on the donor fluorescence quantum yield in the absence of an acceptor, ϕF,D(0), the overlap between the area-normalized donor emission spectrum, F-D(λ), and the acceptor excitation spectrum with extinction coefficient, εA(λ) (in M−1cm−1), at the wavelength λ (in nm), the relative orientation of the dye dipoles captured by the orientation factor, κ2, and the refractive index of the medium, nim, between and around the dyes. It should be noted that, due to the λ4 dependence of the overlap integral, small shifts in the spectra can have large effects on the R0. The following sections describe the factors that influence R0 and the FRET efficiency in more detail.

Extinction coefficient ε

The extinction coefficient of the acceptor dye affects R0 and the expected excitation rate in ALEX/PIE experiments. In the absence of an easy or affordable way to measure this parameter (it requires large amounts of dye for gravimetric analysis or FCS with controlled dilution [Fries et al., 1998]), the experimenter typically relies on the value given by the manufacturer, a value that can at times be unreliable. Alternatively, the extinction coefficient of the dyes may be theoretically assessed via the Strickler and Berg, 1962 equation, when ϕF,D(0) and the fluorescence lifetime are known. Fortunately, ε is not expected to vary much depending on the environment of the fluorophores, since both the ϕF,D(0) and the fluorescence lifetime, in most cases, vary accordingly. Hence, one can conclude that the local environment does not heavily influence the excitation probability (according to the Strickler-Berg equation mentioned above).

Fluorescence quantum yield ϕF

ϕF oftentimes changes upon labeling and can be sensitive to the local environment at the labeling position, to the conformational state of the molecule and to the binding of ligands, substrates or complex partners. Even dyes that are considered relatively insensitive to their local environment have been shown to exhibit a large change in ϕF upon conjugation to nucleic acids or proteins. As an extreme example, the quantum yield of Cy3B ranges from 0.19 to 0.97 at different labeling positions on dsDNA, leading to considerable variation in the value of R0 for the pair Cy3B-ATTO 647N between 54.8 Å and 65.9 Å (Lerner et al., 2018b; Craggs et al., 2019). For dyes of the cyanine family, such as Cy3 and Cy5, or its variants Alexa Fluor 555 and Alexa Fluor 647 (Gebhardt et al., in preparation), ϕF is dependent on the excited-state isomerization, which is influenced by viscosity, steric restriction and (stacking) interactions (Hwang and Myong, 2014; Lerner et al., 2016; Levitus and Ranjit, 2011; Sanborn et al., 2007; White et al., 2006; Widengren et al., 2001). In summary, independent determination of ϕF for different labeling positions is strongly recommended. Notably, nsALEX/PIE and MFD experiments can probe the fluorescence lifetime, and thus directly identify changes in ϕF. Development of standard procedures for measuring or estimating ϕF, for example using an integrating sphere (Gaigalas and Wang, 2008; Pati et al., 2020) or a nanocavity (Chizhik et al., 2013; Chizhik et al., 2011), would benefit the field and should be discussed.

Refractive index nim

The actual index of refraction to be used for calculation of R0 lies somewhere between the index of refraction of an aqueous buffer (1.33) and that for proteins and DNA (~1.5) but the exact value is not known. Robert Clegg recommended using an intermediate value of 1.4, which reduces the maximal error in R0 to ~ 4% (Clegg, 1992). However, different values may be more appropriate depending on the geometry and environment of the fluorophores. To date, the refractive index has received very little attention in the field (Knox and van Amerongen, 2002).

Dye transition dipole orientation factor κ2

This parameter describes the relative orientation of the transition dipole moments of the dyes and strongly depends on dye mobility. Since the dyes’ orientations can change randomly on the time scale of typical FRET events, the mean value of <κ2> = 2/3 is typically taken. This well-known dynamic averaging approximation assumes that the rotational diffusion timescale of a FRET pair is much shorter than the fluorescence lifetime of the donor. However, it may well be that one of the dyes is not freely rotating on this timescale (e.g., it may interact with the microenvironment). An extreme example is a FRET system in which non-canonical fluorescent nucleotides were incorporated into dsDNA. The rigid structure and natural helical twist of the DNA caused the relationship between E and RDA to follow an interesting trend (Ranjit et al., 2009) with E being relatively low around RDAR0, because of κ2 ~ 0 (Wranne et al., 2017). In another smFRET experiment, a DNA molecule was end-labeled with Cy dyes without sulfonic acids groups (Cy3 and Cy5), which have a tendency to stack onto bases at the DNA termini (Iqbal et al., 2008; Ouellet et al., 2011), and the influence of orientational effects on the FRET efficiency was measured. Although an influence of the orientation could be detected, the data showed that orientational effects average-out quite well in most realistic cases (Iqbal et al., 2008). A method to estimate the lower and upper bounds for 2> from the donor and acceptor time-resolved anisotropies was proposed in the 1970s (Dale et al., 1979; van der Meer, 2002). In smFRET measurements using the polarization-resolved MFD modality, information on the donor and acceptor fluorescence intensities, lifetimes, and anisotropies (Schaffer et al., 1999) are collected simultaneously and fluorescence anisotropy decays of different single-molecule sub-populations can be used to assess the 2> uncertainty per conformational state (Ivanov et al., 2009; Kudryavtsev et al., 2012; Sindbert et al., 2011). It is noteworthy to mention that the majority of fluorophores used as donor and acceptor dyes in smFRET have a mono-exponential fluorescence decay and, hence, have one major emission dipole. In this case, the estimation of 2> depends on the orientation of these single transition dipole moments. It has been proposed that the assumption of 2> = 2/3 would carry much less uncertainty when the fluorescence signal is emanating from more than one emission dipole, yielding multi-exponential decays (Haas et al., 1978). This is an intriguing idea that could provide a realistic estimation for κ2 and thus help simplify the transformation of FRET efficiencies into inter-dye distances. For a review on the dependence of FRET on κ2, the reader is referred to (van der Meer, 2002). Finally, we note that there are several routines recommended by experienced members of the community to determine R0 accurately. Which approach is the most optimal is still under discussion.

Dye models

The Förster equation (Equation 1) allows the extraction of a distance directly from a FRET efficiency measurement. This distance directly corresponds only to the separation of the FRET fluorophores when the positions of the donor and the acceptor molecules are constant, the dye’s orientations are rapidly averaged and their microenvironment is known. Strictly speaking, this is never the case, even for a stable conformation of the labeled macromolecule, since dye molecules are typically attached to the macromolecules via flexible linkers, and the labeled macromolecule usually restricts the volume accessible to the fluorophore (Best et al., 2007; Hellenkamp et al., 2018a; Ingargiola et al., 2018b). In addition, diffusion of the dyes while the donor is in the excited state can also influence the measured FRET efficiency (Ingargiola et al., 2018b). When the FRET rate is not too high (leading to a FRET efficiency of E<0.8) and the dyes do not interact with the protein surface, deviations due to dye dynamics are usually negligible (Hellenkamp et al., 2018a; Hellenkamp et al., 2017; Kalinin et al., 2012).

Various groups have developed detailed dye models that account for the translational and rotational flexibility of the dyes and thus allow a more accurate description of the actual distance FRET measures (Figure 5D) (Beckers et al., 2015; Craggs and Kapanidis, 2012; Dimura et al., 2016; Haas et al., 1978; Kalinin et al., 2012; Muschielok et al., 2008; Schuler et al., 2020; Sindbert et al., 2011) and to test them experimentally (Hellenkamp et al., 2018a; Nagy et al., 2018; Peulen et al., 2017; Wozniak et al., 2008). For any given FRET efficiency, different dye models will lead to slightly different extracted RDA distributions (deviations ≤ 5%). Choosing an appropriate model is therefore important for the accurate determination of RDA.

Dye models for FRET.

(A) The different kinetic averaging regimes for rotation and diffusion are shown schematically. In the dynamic averaging regime, rotation and diffusion happen on a timescale faster than the FRET process. In the isotropic averaging regime, translation is slower than the FRET process, but rotation is fast. The static average applies if both rotation and diffusion are slow compared to the rate of energy transfer. For most experimental situations, the isotropic average is most appropriate. (2008 National Academy of Sciences. Reproduced from Wozniak et al., 2008. Further reproduction of this panel would need permission from the copyright holder.) (B) A schematic of a donor/acceptor fluorophore pair (Alexa488, Alexa647) attached to the protein Atlastin-1. (C) The accessible volume that the acceptor fluorophore can explore is determined from a molecular dynamics simulation. (D) Different coarse-grained dye models are used in the community to describe the three-dimensional dye density ρDye (see main text). (Panels B, C, and D were reproduced from Dimura et al., 2016, Current Opinion in Structural Biology with permission, published under the Creative Commons Attribution 4.0 International Public License (CC BY 4.0; https://creativecommons.org/licenses/by/4.0/).)

© 2008, National Academy of Sciences. Panel A was originally published as Figure S2 in Wozniak et al., 2008. Further reproduction of this panel would need permission from the copyright holder.

© 2016, Dimura et al. Panels B, C and D were originally published as Figure 1a-d in Dimura et al., 2016, published under the Creative Commons Attribution 4.0 International Public License.

Changes in RDA and the relative orientation of the dyes can occur on many different time scales, including the excited state lifetime, the interphoton time, and the photon burst duration. Averaging over these different distances and orientations is complicated due to the inherent non-linearity of the energy transfer process. However, there are exact analyses that describe the photon statistics in smFRET experiments (Antonik et al., 2006; Gopich and Szabo, 2005a; Nir et al., 2006).

Since the dye linker lengths of the typical dyes used in smFRET experiments are long (ranging typically between 10 and 15 atoms, Figure 5B), translational and rotational diffusion of the dyes within the accessible volumes constrained by their linkers and the macromolecules to which they are conjugated lead to considerable changes in RDA. Such dynamics can occur on timescales comparable to the fluorescence lifetime, which leads to changes in RDA, from the moment of donor excitation to the moment of donor de-excitation. This process leads to the well-documented phenomenon termed diffusion-enhanced FRET (Beechem and Haas, 1989; Haas and Steinberg, 1984; Orevi et al., 2014), where the measured FRET efficiencies are higher than expected from a static distribution of RDA due to the increase in the probability for FRET to occur when RDA shortens while the donor is in the excited state (Eilert et al., 2018; Ingargiola et al., 2018b).

This phenomenon has been treated by incorporating both rotational and translational diffusion of the fluorophores as fluctuations in RDA inside a potential well of the reaction coordinate RDA (Dingfelder et al., 2018; Haas and Steinberg, 1984; Ingargiola et al., 2018b). Similarly, rotational motions lead to changes in the relative orientation of the donor and the acceptor molecules and therefore to changes in R0 via κ2. To this end, a complete kinetic theory treating both rotational and translational diffusion has been developed (Eilert et al., 2018). In many cases, a dynamic rotation - static translation model can be used (i.e., krotation>>kFRET>>ktranslation) (Figure 5A). Interestingly, Monte-Carlo simulations show that this often-applied simplification can lead to errors in RDA (Hellenkamp et al., 2018a). The magnitude of the uncertainty depends on the donor fluorescence lifetime, the FRET efficiency, and the dye molecules’ diffusion constants and rotational correlation times. So far, no major disagreement of the dynamic rotation – static translation model with experimental data has been reported, thus supporting the use of the isotropic average of <κ2> = 2/3.

To obtain atomistic insights into the behavior of dyes on biomolecules, molecular dynamics simulations have been explored (Best et al., 2007; Deplazes et al., 2011; Spiegel et al., 2016; Girodat et al., 2020; Grotz et al., 2018; Hoefling et al., 2011; Reinartz et al., 2018; Shoura et al., 2014). By simulating the whole system, including the fluorophores, information is obtained about the accessible volume of the fluorophore, its potential interactions with the biomolecular surface and the dynamics of the system. The results of such simulations crucially depend on the parameterization (force field) of the dyes. Different parameter sets have been reported for commonly used dyes and validated against experimental data (Best et al., 2015; Graen et al., 2014; Schepers and Gohlke, 2020; Shaw et al., 2020), but a consensus on the optimal parameterization has not yet been reached.

Structural modeling

By accounting for various uncertainties described in the section Inter-dye distances, precise distances can be calculated from FRET efficiencies. This enables the application of FRET for FRET-based structural studies, which are particularly promising for studying the conformations of large, heterogeneous, flexible, and dynamic biomolecules and their complexes (Brunger et al., 2011; Craggs et al., 2019; Filius et al., 2020; Hellenkamp et al., 2017; Holmstrom et al., 2018; Kilic et al., 2018; Muschielok et al., 2008; Nagy et al., 2015; Sanabria et al., 2020; Treutlein et al., 2012; Yanez Orozco et al., 2018). Such systems are notoriously difficult to study with classical structural biology methods. For structure determination, there are further steps that need to be taken.

Pipeline

Different approaches have been used to derive structural models from FRET distance restraints. The general pipeline consists of the following steps:

  • preparation and measurement of multiple donor-acceptor labeled variants with different labeling positions,

  • control experiments to assess the activity after labeling or immobilization, the photophysics of the probes, and the rotational freedom of the dyes,

  • the non-trivial transformation from proximity ratios (uncorrected E values) to absolute FRET efficiencies of the different conformational subpopulations, to inter-dye distance information (or equilibrium distance distributions),

  • relating the inter-dye distances to the structure by an appropriate dye model, and

  • assessing the quality of the structural model.

The FRET information alone is insufficient to generate an atomistic model de novo. FRET-restrained structural modeling thus relies on prior structural knowledge, from which novel structural models are generated. Different approaches have been used, each with specific advantages and limitations (Figure 6).

Approaches to integrative modeling using FRET restraints.

(A) Rigid bodies, representing either different domains of a protein or different proteins within a complex, are arranged by translation and rotation to satisfy the FRET restraints. Adapted from Figure 1 of Hellenkamp et al., 2017. (B) The structure that best satisfies all FRET restraints is selected from a prior ensemble. Adapted from Figure 2 of Dimura et al., 2020. (C) Using triangulation, the most likely dye position is estimated from the FRET distances with respect to known reference positions on the structure. (Reproduced from Andrecka et al., 2009, Nucleic Acid Research with permission, published under the Creative Commons Attribution-NonCommercial 4.0 International Public License (CC BY NC 4.0; https://creativecommons.org/licenses/by-nc/4.0). Further reproduction of this panel would need permission from the copyright holder.) (D) Starting from a known structure, a molecular dynamics simulation is guided using the FRET information by applying forces based on the FRET distances. (Reproduced from Dimura et al., 2020. Further reproduction of this panel would need permission from the copyright holder.)

© 2009, Andrecka et al. Panel C was originally published as Figure 2A in Andrecka et al., 2009, published under the Creative Commons Attribution-NonCommercial 4.0 International Public License.

© 2020, Dimura et al. Panel D was originally published as Figure 2A in Dimura et al., 2020. Published under the Creative Commons Attribution 4.0 International Public License.

Different laboratories use different approaches to analyze their experimental results in context-dependent manners and the steps necessary for the conversion from FRET information to inter-dye distance information are still under debate. Importantly, the overall uncertainties in inter-dye distances need to be determined from the uncertainty in R0, the dye model and experimental precision, and to be incorporated into the integrative structural modeling (Dimura et al., 2016; Hellenkamp et al., 2018a; Hellenkamp et al., 2017; Kalinin et al., 2012; Muschielok et al., 2008; Muschielok and Michaelis, 2011; Peulen et al., 2017). As a result, FRET restraints yielded structural models that showed excellent agreement with existing models in benchmark studies (Kalinin et al., 2012; Mekler et al., 2002), resolved the structure of flexible parts and short-lived excited conformational states that were inaccessible in crystallographic studies (Andrecka et al., 2009; Sanabria et al., 2020) and quantified the conformational flexibility of crystallographic states (Hellenkamp et al., 2017). To increase the success of FRET-derived structural models, a protocol has recently been proposed to determine the most informative labeling positions, the number of labeling positions needed to resolve a given conformation and the accuracy that can be expected for the FRET-derived structural models (Dimura et al., 2020).

We should also be aware of the expected uncertainty for the determined structural model, which can be computed from the number of measurements, their average quality and the properties of the underlying computational method. Besides estimating the precision (Dimura et al., 2020), it is useful to introduce a quantitative quality estimate, χn2, for judging the accuracy of FRET-derived structure models based on a cross-validation statistical method in the spirit of Rfree used in X-ray crystallography (Brünger, 1992). Such accuracy estimates are essential for the quality control of I/H structure models, especially those deposited in the PDB-Dev. Dimura et al., 2020 pointed out that the complexity arising from different experiments could be mitigated by using Bayesian hierarchical data processing frameworks, which abstracts the experimental data and propagates the information and uncertainties to enable structural modeling at higher precision and accuracy. Combining the smFRET information with additional constraints provided by complementary methods has the potential to further improve the accuracy of the obtained I/H structural models (Berman et al., 2019).

Verification of steps in the workflow using simulations

At various points along the workflow, simulations can be extremely useful. They can be used to gain a better understanding of the analysis procedures (e.g., to test the impact of specific burst search and selection parameters) (Hagai and Lerner, 2019) or to determine to what extent a particular data analysis procedure is capable of extracting results reliably (e.g., distinguishing between one or two subpopulations) (Blanco and Walter, 2010; Chen et al., 2016). Having a ‘ground truth’ to compare to is also very helpful when developing new analysis methods before applying it to real data. This question is relevant, especially when other common analysis procedures yield different results. Of course, this practice should not replace control measurements and analyses with an established system for validation of a given method, although numerical simulations can give an economical and initial guidance. Using prior knowledge about the measured system and the experimental setup, one can perform simulations mimicking real measurements and then check the results of different analysis procedures. In this way, it is possible to determine which analysis procedure is most promising in a given context. Another insightful aspect that simulations can provide is a consistency check of the data. One can simulate smFRET data assuming the number of states and/or kinetic rates derived from the experiments to see how well the model describes the measured data.

Open science

One of the cornerstones of the scientific method is the ability to reproduce experimental results. As experiments become more sophisticated, a clear description of the experiments is crucial. Recent trends toward Open Science practices call for full transparency of the scientific process, as has been formulated in the FAIR principle that data should be ‘Findable, Accessible, Interoperable, and Re-usable’ (Wilkinson et al., 2016). To this end, the procedures taken to acquire, analyze, and interpret experimental data should be provided. That includes describing each step, the reasons for taking the step and the information associated with the step. To ensure that the analysis remains transparent and tractable, code should generally be openly available and all parameters and settings used in the analysis should be stored.

Funding agencies embracing this philosophy (e.g., https://datascience.nih.gov/strategicplan) expect grantees to publish in open-access (OA) journals (and pay for the corresponding OA fees) or deposit manuscripts on preprint servers (e.g., Pubmed Central, arXiv, bioRxiv, ChemRxiv, medRxiv), and deposit data (sometimes also raw data) in repositories (e.g., Zenodo, the Dryad Digital Repository, FigShare) as well as analyses codes (e.g., GitHub). Open science disseminates knowledge by freely sharing results and the tools developed by independent scientists or teams working as part of a collaborative network. We would like to see the FRET community embrace and be committed to open science. Some tools are already in place, while others still need to be developed to make it easier to communicate the continuously growing knowledge and experience present in the FRET community.

Intellectual property and software licenses

There is obviously some tension between the precepts of open science and requirements imposed by some intellectual property (IP) policies. IP rights, including patent laws, were put in place to promote the development of science and technology for the benefit of society by allowing those developing intellectual property to retain the rights for the IP they developed. In fact, in some sense, patents were the first form of open access publication, only with a restrictive license for reuse. We do not oppose intellectual property rights, but given the developmental stage the FRET field is currently in, we support the disclosure of methods, data, and software. For the advancement of the field, other groups must be able to reproduce the analyses of existing data, extend upon them and, if needed, be able to reproduce the experiments. The acquisition and analysis must be modifiable and extendable in agreement with the license chosen by the data or software creator. This license should be set as liberal as possible, taking into account the IP considerations mentioned above, but also encourage recognition of the considerable effort invested in producing successful protocols, designs, data, or software. Ultimately, if practiced fairly, open science should entice everyone, including commercial vendors, to adopt and contribute to community-defined file formats, provide free file-conversion codes, and open their analysis tools for scrutiny by the community.

Proper documentation of data analysis practices

By making analysis codes and protocols freely available, we hope to stimulate the acceptance, utilization, and exchange of new methods and tools. It is true that there already exist a large number of open-access programs that offer a large variety of analysis procedures for single-molecule photon trajectories (free-diffusion smFRET) and single-molecule videos (immobilized smFRET) data (see Table 1).

Table 1
List of available software for smFRET analysis.
SoftwareTypeDescriptionURL
ALiXConfocalALiX is developed for basic research on diffusing two-color smFRET in single or multiple spot geometries (Ingargiola et al., 2017).https://sites.google.com/a/g.ucla.edu/alix/
FreticaConfocalFretica, a Mathematica package with a backend written in C++, is a user-extendable toolbox that supports MFD, PIE/ALEX, PCH (Huang et al., 2004; Müller et al., 2000), FIDA (Gopich and Szabo, 2005b; Kask et al., 1999), PDA (Antonik et al., 2006; Ernst et al., 2020), recurrence analysis (Hoffmann et al., 2011), fluorescence lifetime fitting, FLIM, FCS, FLCS (Arbour and Enderlein, 2010; Dertinger et al., 2007), dual-focus FCS, nsFCS (Nettels et al., 2007; Schuler and Hofmann, 2013), maximum likelihood estimation from photon-by-photon (Gopich and Szabo, 2009) and binned trajectories, simulation of confocal experiments and more (Nettels and Schuler, 2020https://schuler.bioc.uzh.ch/programs/).https://schuler.bioc.uzh.ch/programs/
FRET_3colorCWConfocalC++ and MATLAB GUI-based CPU-GPU co-parallelization software package for an enhanced maximum likelihood analysis of two- and three-color fluorescence photon trajectories generated by continuous-wave donor excitation (Yoo et al., 2020).https://github.com/hoisunglab/FRET_3colorCW
gSMFRETdaConfocalgSMFRETda is a GPU-capable program for Monte Carlo simulations of PDA. It can sample dwell time and other parameters in fine grids, thus allowing the analysis of rapid dynamic interconversions (Liu et al., 2019).https://github.com/liu-kan/gSMFRETda
H2MMConfocalH2MM is a maximum likelihood estimation algorithm for a photon-by-photon analysis of smFRET experiments (Pirchi et al., 2016).http://pubs.acs.org/doi/suppl/10.1021/acs.jpcb.6b10726/suppl_file/jp6b10726_si_002.zip
MFD Spectroscopy and ImagingConfocalA modular software package for confocal fluorescence spectroscopy and imaging experiments using multiparameter fluorescence detection (MFD) with all tools (FCS, fFCS, PDA, seTCSPC, trace analysis, 2D simulation of MFD diagrams, Burbulator) and multiparameter fluorescence image spectroscopy (MFIS) (Antonik et al., 2006; Felekyan et al., 2005; Kühnemuth and Seidel, 2001; Widengren et al., 2006Felekyan et al., 2020).https://www.mpc.hhu.de/software/3-software-package-for-mfd-fcs-and-mfis
OpenSMFSConfocalA collection of tools (Ingargiola et al., 2016b) for solution-based single-molecule fluorescence spectroscopy, including smFRET, FCS, MC-DEPI (Ingargiola et al., 2018b).https://github.com/OpenSMFS
smfBoxConfocalA confocal smFRET platform, providing build instructions and open-source acquisition software (Ambrose et al., 2020).https://craggslab.github.io/smfBox/
rFRETConfocal
Imaging
rFRET is a comprehensive, MATLAB-based program for analyzing ratiometric microscopic FRET experiments (Nagy et al., 2016).https://peternagy.webs.com/fret#rfret
ChiSurfConfocal
Imaging
Ensemble
ChiSurf is a fluorescence analysis platform for the analysis of time-resolved fluorescence decays (Peulen et al., 2017).https://github.com/Fluorescence-Tools/ChiSurf/wiki
PAM - PIE Analysis with MATLABConfocal
Imaging
Ensemble
PAM (PIE analysis with MATLAB) is a GUI-based software package for the analysis of fluorescence experiments and supports a large number of analysis methods ranging from single-molecule methods to imaging (Schrimpf et al., 2018Barth et al., 2020).RRID:SCR_020966, https://www.cup.uni-muenchen.de/pc/lamb/software/pam.html
AutoSiMTIRFAutoSiM is a deep-learning developed MATLAB program for automatically selecting and sorting smFRET traces (Li et al., 2020a).https://doi.org/10.7302/ck2m-qf69
BIASDTIRFBIASD uses Bayesian inference to infer transition rates that are more than three orders of magnitude larger than the acquisition rate of the experimental smFRET data (Kinz-Thompson and Gonzalez, 2018).http://github.com/ckinzthompson/biasd
DeepFRETTIRFsmFRET software based on deep-learning for automatic trace selection and classification. It includes all common features: image analysis, background-corrected trace-extraction, hidden Markov analysis, correction factor application, and data visualization (Thomsen et al., 2020).https://github.com/hatzakislab/DeepFRET-GUI
ebFRETTIRFebFRET performs combined analysis on multiple smFRET time-series to learn a set of rates and states (van de Meent et al., 2014).https://ebfret.github.io/
FRETboardTIRFsmFRET data preprocessing and analysis using algorithms of choice and user supervision. Also offered as a web-based user interface (de Lannoy et al., 2020).https://github.com/cvdelannoy/FRETboard
HaMMyTIRFsmFRET analysis and hidden Markov modeling (McKinney et al., 2006).http://ha.med.jhmi.edu/resources/
hFRETTIRFhFRET uses variational Bayesian inference to estimate the parameters of a hierarchical hidden Markov model, thereby enabling robust identification and characterization of kinetic heterogeneity (Hon and Gonzalez, 2019).https://github.com/GonzalezBiophysicsLab/hFRET
iSMSTIRFiSMS is a user-interfaced software package for smFRET data analysis. It includes extraction of time-traces from movies, traces grouping/selection tools according to defined criteria, application of corrections, data visualization and analysis with hidden Markov modeling, and import/export possibilities in different formats for data-sharing (Preus et al., 2015Preuss et al., 2020).http://isms.au.dk/
MASH-FRETTIRFMASH-FRET is a MATLAB-based software package for the simulation (Börner et al., 2018) and analysis of single-molecule FRET videos and trajectories (video processing [Hadzic et al., 2016], histogram analysis [König et al., 2013], and transitions analysis [Hadzic et al., 2018; König et al., 2013]).https://rna-fretools.github.io/MASH-FRET/
miCUBETIRFTIRF smFRET platform, providing detailed build instructions and open-source acquisition software (Martens et al., 2019).https://hohlbeinlab.github.io/miCube/index.html
SMACKSTIRFSMACKS (single-molecule analysis of complex kinetic sequences) is a maximum-likelihood approach to extract kinetic rate models from noisy single-molecule data (Schmid et al., 2016).https://www.singlemolecule.uni-freiburg.de/software/smacks
smCameraTIRFsmFRET data acquisition (Windows. exe) and analysis (IDL, MATLAB) with example data (Roy et al., 2008).http://ha.med.jhmi.edu/resources/
SPARTANTIRFAutomated analysis of smFRET multiple single-molecule recordings. It includes extraction of traces from movies, trace selection according to defined criteria, application of corrections, hidden Markov modeling, simulations, and data visualization (Juette et al., 2016).https://www.scottcblanchardlab.com/software
STaSITIRFSTaSI uses the Student’s t-test and groups the segments into states by hierarchical clustering (Shuang et al., 2014).https://github.com/LandesLab/STaSI
TwoToneTIRFA TIRF-FRET analysis package for the automatic analysis of single-molecule FRET movies (Holden et al., 2011).https://groups.physics.ox.ac.uk/genemachines/group/Main.Software.html
vbFRETTIRFvbFRET uses variational Bayesian inference to learn hidden Markov models from individual, smFRET time trajectories (Bronson et al., 2009).http://www.columbia.edu/cu/chemistry/groups/gonzalez/software.html
Fast NPSModelingA nano-positioning system for macromolecular structural analysis (Eilert et al., 2017).http://dx.doi.org/10.17632/7ztzj63r68.1
FluordynamicsModelingFluordynamics is a PyMOL plugin to label biomolecules with organic fluorophores for all-atom molecular dynamics simulations. It builds on AMBERDYES (Schepers and Gohlke, 2020) and extends the force field to common nucleic acid linker chemistries (Steffen et al., 2016).https://github.com/RNA-FRETools/fluordynamics
FPSModelingA toolkit for FRET restrained modeling of biomolecules and their complexes for quantitative applications in structural biology (Kalinin et al., 2012).https://www.mpc.hhu.de/software/1-fret-positioning-and-screening-fps
FRETrajModelingFRETraj is a Python API to the LabelLib package, which integrates into PyMOL to interactively calculate accessible-contact volumes and predict FRET efficiencies (Steffen et al., 2016).https://github.com/RNA-FRETools/fretraj
FRETrest in Amber20ModelingFRETrest is a set of helper scripts for generating FRET-restraints for Molecular Dynamics (MD) simulations performed with the AMBER Software Suite (Dimura et al., 2020).http://ambermd.org/doc12/Amber20.pdf
LabelLibModelingLabelLib is a C++ library for the simulation of the accessible volume (AV) of small probes flexibly coupled to biomolecules (Dimura et al., 2016; Kalinin et al., 2012).https://github.com/Fluorescence-Tools/LabelLib

To further improve the inter-operation between methods and analysis and to establish convenient documentation protocols, it is essential to work in an open multivalent environment. For this goal, the use of browser-based software such as ‘Jupyter notebooks’ and/or other available workspaces may serve as a convenient platform. Such workspaces provide an interactive scripting environment by combining formatted ‘rich’ text with well-commented code commands as well as code outputs (e.g., figures, tables, comments, equations) and explanations in a single web-based document. Such web-based workspace environments support several programming languages, including Python, R, C++, and, to some extent, MATLAB and Mathematica. Practitioners using this environment can then easily read, distribute, re-run, check, and modify the code. Software engineering approaches in scientific software usually include version control, code review, unit testing, continuous integration, and auto-generation of HTML manuals. In the next step, Jupyter notebooks or similar workspaces can also help newcomers perform complex analyses already in the web-based environment with minimal adaptation efforts, which will accelerate the dissemination of new analyses. Indeed, well-documented, easy-to-use notebooks have been provided by various groups (Ambrose et al., 2020; Ingargiola et al., 2016b; Ingargiola et al., 2016a; Lerner, 2020; Lerner, 2019) (e.g., at https://github.com/tritemio/FRETBursts or https://craggslab.github.io/smfBox/).

Although the notebook approach offers advantages to experienced users and software developers, it might be difficult for many end-users to adapt to the script-based workflow. For those users, it may be more convenient to use the established and tested algorithms embedded in a graphical user interface (GUI). Indeed, there is a large variety of user-friendly software available (compiled in Table 1). To further increase the ease of use, the FRETboard software aims to make the underlying analysis algorithms of other packages available through a single web-based GUI (de Lannoy et al., 2020). As it can be used in a browser through a remote web server, this would allow any user to freely experiment with different analysis methods without the need for software installation or heavy computational resources.

As the first step toward FAIR-compliant analysis practices, we propose establishing a software library that contains tested and proven algorithms for the analysis of fluorescence experiments, which will assist in their efficient distribution and implementation in existing workflows. Such efforts have already been initiated in the FRETbursts software package (Ingargiola et al., 2016b), and a GitHub group has been established at https://github.com/Fluorescence-Tools to collect software packages and connect software developers. Establishing a community-wide working group of ‘Analysis software for FRET’ would be an important step in organizing and moderating this process.

Standard file format

To expedite the exchange of data between different groups and testing of different analysis methods, it would be beneficial to have a minimal number of file formats, and to avoid the multiplication of ad hoc formats developed independently. In fact, the absence of a standard file format and supporting documents has caused issues within individual labs with respect to long-term data storage and re-analysis. Online data deposition in well-documented file formats would therefore save a lot of headaches for many laboratories. Even if standard file formats are carefully designed or developed, it will be inevitable to modify the existing formats or introduce new formats in the future. The list of standard formats should be regularly updated, and the analysis codes should also be kept current and properly maintained to guarantee their compatibility with the new standard formats.

The raw experimental data should be supplied in a universal data file format that can be easily read and scrutinized (Figure 7 left). Ideally, the file should store both raw data and sufficient metadata to specify the measurement, setup, and sample. The metadata should be stored in a human-readable text-based format, while space-efficient storage of the raw photon data should be ensured by lossless compression. There are currently many different file formats for smFRET data, developed by different research groups and companies. In general, such files are hard to access for other laboratories and they are not guaranteed to be perennial, which poses an additional challenge to the community. To promote the adaptation of new file formats, conversion tools for older file formats should be provided so that future software codes can focus on handling one (or at least only a very few) common file formats, such as the phconvert suite of notebooks for transforming many file formats into Photon-HDF5 (see http://photon-hdf5.github.io/phconvert/). Reaching such a consensus is possible, as has been demonstrated by the acceptance of a single file format for the deposition of NMR data (Ulrich et al., 2019).

Concept for data storage following the FAIR principle.

All essential information should be contained in two data files, one for raw data and a second with the essential information regarding the associated analysis. (Left) Structure of proposed data file formats containing confocal or TIRF raw data in a time-tagged, TT, or time-tagged time-resolved, TTTR, format together with sample- and experiment-specific metadata (for details, see Figure 8). The measurement specifications are needed by the reading routine to reconstruct the photon trace from the stored data, for example, for timing in the TT format (sampling frequencies expressed as time bins or frame rates) or in the TTTR format (laser repetition rates, time binning in time-correlated single-photon counting). Moreover, the detector that measured the signal is noted (detector #) along with the detection time with a given time resolution. For representing the detection time of single photons in time-resolved smFRET studies, the TTTR format is used where the time corresponds to the sum of the macro time and the nano time (upper left panel). The macro time comprises multiple cycles of excitation laser pulses (blue vertical lines) and the nano time is determined by time-correlated single-photon counting with picosecond resolution. The TTTR format is the most compact data format for single-molecule fluorescence data for detection times with picosecond time resolution and macro-times of hours (Felekyan et al., 2005). The representation of the photon detection of intensity-based and imaging-based smFRET studies is in the TT format, where the macro-time comprises multiples of the external clock pulse or readout cycles, where single or several photons were detected. The stack of TIFF images acquired in TIRF measurements (lower left panel) is transformed into the TT format for analyzing photon time traces for selected spots. (Middle) For the corresponding data file, a metadata system as implemented in the Photon-HDF5 file format (Ingargiola et al., 2016a) is suggested. (Right) The analysis file should contain the determined parameters obtained by a quantitative analysis together with analysis metadata that assure evaluation, reproducibility, and re-usability of the analysis. The processed data should be documented as outlined in Figure 8D.

File formats for (time-correlated) photon counting with point-detector data (confocal modality)

Several formats have been used and reported for time-correlated single-photon counting data (Brand et al., 1997; Brooks Shera et al., 1990; Eggeling et al., 2001; Felekyan et al., 2005; Rigler et al., 1993; Schaffer et al., 1999; Tellinghuisen et al., 1994; Wahl et al., 2008; Widengren et al., 2006). One example is the time-tagged time-resolved (TTTR) file format given in the left panel of Figure 7. Because of its compactness, this file format has been widely adopted by commercial companies (e.g., Becker and Hickl and PicoQuant) providing TCSPC electronics and point detectors for solution/imaging studies.

The basic formats have been extended in the Photon-HDF5 (Ingargiola et al., 2016a) file format that connects rich metadata with the raw photon information in a single, space-efficient format suitable for sharing and long-term data archival. Moreover, several software programs exist that can easily transform raw data files to the Photon-HDF5 format. Once enough metadata is available in the community, the relative importance of particular metadata entries on the resulting FRET data can be assessed.

File format for camera-based acquisition (wide-field/TIRF modality)

Camera-based data is acquired as a stack of images (e.g., TIFF). To extract time trajectories, several steps are needed to yield the time-binned fluorescence intensity (e.g., spot identification; donor and acceptor spot registration; thresholding; background subtraction, etc.). A file format for TIRF-based smFRET (immobilized) measurements has been proposed (Greenfeld et al., 2015). Alternatively, human-readable plain text files with an agreed format and greatly downsized from the raw TIFF stacks can also work well for TIRF-based smFRET traces. Binary file formats for efficient data compression have also been implemented (Juette et al., 2016).

Exchange file format for processed data

In addition to standardized raw file formats, we recommend defining exchange file formats for different levels of processed data (Figure 7, middle panel). This will allow researchers to establish flexible and modular workflows spanning different software packages and facilitate the adoption of novel analysis approaches. The deposition of processed data in agreed-upon file formats also ensures that published data can be re-used at a later time point, for example, for more elaborate structural modeling approaches (Köfinger et al., 2019). For the storage of processed data, we believe that it is important to retain the connection to the raw data by including the relevant metadata. For example, corrected FRET efficiency histograms should be deposited together with the raw signal intensities in the donor and acceptor channels, the background intensities and the calibration factors.

Repositories and data bank for FRET data and models

Metadata for FRET experiments

To ensure that the reported results are reproducible, raw data must be sufficiently annotated. One important source of inspiration for developing recommendations and standards for the FRET community comes from the world-wide Protein Data Bank (wwPDB) (Berman et al., 2003; Young and Westbrook, 2019). Following the standards of the wwPDB, it is recommended to provide additional information to the models as outlined in Figure 8.

Disseminating the results of smFRET studies.

Recommended categories for data and method-specific information (metadata), which are needed for documentation of smFRET studies where the authors want to archive their obtained kinetic/structural models. (A) General information. (B) Information on the sample and FRET-specific properties. (C) Information on the experiment and data acquisition. (D) Information on processing and analysis procedures. (E) Information on the interpretation of the data and the final kinetic network or structural model.

Archiving FRET experiments

In this respect, we strongly encourage the publication of datasets with detailed descriptions of the acquisition and analysis in scientific journals. Alternatively, or in addition, structural information and raw datasets should immediately be deposited in repositories provided by the publisher or generalist repositories (Table 2) with a Digital Object Identifier or link for access and citation. Online and public repositories also act as a form of data and knowledge backup, which is difficult to achieve and maintain at the scale of a single laboratory. Comparing a repository with a database, it is obvious that a database has significant advantages for the scientific community. While both archive forms provide open access to the data for all users, only a database can fulfill the quality criteria for safe usage of the deposited data by (1) creating standard definitions for the experimental data used for determining kinetic and/or structural models and their features; (2) developing methods to collect the minimum amount of required information for curation and validation of models and data; and (3) building the infrastructure for acquiring, archiving and disseminating the models and the data. To conclude, a database assures appropriate documentation with a defined format and quality control of the results.

Table 2
A list of repositories for citable data storage.
Repository nameURLSize limitsFee/Costs
Dryad Digital Repositoryhttps://datadryad.org/stashNo$120 USD for first 20 GB, and $50 USD for each additional 10 GB
figsharehttps://figshare.com/1 TB per datasetIt varies with publishers
Harvard Dataversehttps://dataverse.harvard.edu/2.5 GB per file, 10 GB per datasetFree
Open Science Frameworkhttps://osf.io/5 GB per file, multiple files can be uploadedFree
Zenodohttps://zenodo.org/50 GB per datasetDonation
Mendeley Datahttps://data.mendeley.com/10 GB per datasetFree

Given the importance of integrative structures for advancing life sciences and the significant world-wide investment made to determine them, the wwPDB has proposed a governance structure for federating archives containing structural models (e.g., PDB, Electron Microscopy Data Bank, EMDB Tagari et al., 2002) or experimental data. Moreover, for integrative modeling integrative/hybrid modeling structures (Sali et al., 2015), PDB-Dev (Berman et al., 2019; Vallat et al., 2018) was established as an associated prototype system for archiving multi-state and ensemble structures on multiple spatial scales. In addition, associated kinetic models can be deposited.

To deposit structural models and kinetic networks obtained using FRET experiments in PDB-Dev, a dictionary for FRET data is under development (https://github.com/ihmwg/FLR-dictionaryVallat et al., 2020), serving as a method-specific extension to the existing integrative/hybrid modeling (IHM) dictionary that contains the data categories described in Figure 8. Presently, three integrative FRET-assisted structural models can be found on PDB-Dev.

In addition to archiving the models themselves, all relevant experimental data and metadata should be archived in a technique-associated data bank similar to the Biological Magnetic Resonance Data Bank, BMRB (Ulrich et al., 2019; Ulrich et al., 2008). Reaching such a consensus is possible, as has been demonstrated by the acceptance of a single file format for the deposition of NMR data (Ulrich et al., 2019). In the future, federated resources for other biophysical techniques are expected to align with the structural model archives of the wwPDB for participating in data exchange. Thus, it would be desirable to establish a federated data bank for archiving FRET and, more generally, fluorescence data that could be referred to as the Fluorescence Data Bank (FLDB).

Community actions to bring FRET scientists together

To better achieve a consensus on the current and the future directions of the smFRET community, an open forum is needed where the current issues, needs, and desires could be discussed. We propose the following tools to organize the community around standardization efforts and open science practices. Some of these tools have already been put in place.

Community website as a central hub

A website for the FRET community has been established at https://www.fret.community. The community is open to everybody and registered members can populate their user profiles with additional information such as a description of their scientific interests or a list of key-publications. Besides providing regular updates on the activities within the community, the website also provides resources such as a curated list of software packages (see Table 1) and offers a discussion platform through an integrated forum. The website serves as a platform for ongoing discussions, announcements of accepted relevant papers, notifications about upcoming meetings, workshops, competitions and other activities that might be relevant to the community. An advisory board, elected by the community, moderates the website. One can also envision adding an educational section, much like the popular website for general microscopy education (https://micro.magnet.fsu.edu).

Listserv

To facilitate the dissemination of important information to the FRET community, an electronic mailing list (Listserv) has been established. In order to subscribe to it, smFRET practitioners are requested to register (free of charge) using the following link: https://www.fret.community/register. The members will be informed through the email list about ongoing activities and developments within the community, such as experimental or computational challenges, key publications in the fields, and workshops or meetings.

Server and repository

A repository will be established, which will be accessible through the community website, to host a collection of software packages and facilitate the community-driven joint development of analysis tools. The repository will contain dedicated sections for acquisition software, raw data, analysis codes, analyzed data files, and file conversion utilities. In order to deposit code in the repository, guidelines for the required documentation will be provided.

The concept of the repository is to support open science and transparency. Anyone registered on the website will be able to access raw data, and analyze and compare performances of the various analysis codes. Moreover, the codes can be updated and expanded (while keeping original versions) by anyone. In this way, improvements and enhancements can be implemented and tested. In that context, it is important to mention that such a repository can also serve the purpose of source data deposition, nowadays required by many scientific journals.

Participation in CASP(-like) competitions

Critical Assessment of protein Structure Prediction (CASP, http://predictioncenter.org/) is a grass-roots effort for predicting a three-dimensional protein structure from its amino acid sequence. CASP has been run, since 1994, as a double-blind competition. It provides research groups with an opportunity to test their structure prediction methods objectively. CASP has been exploring modeling methods based, in part, on sparse experimental data, including data from SAXS, NMR, crosslinking, and FRET. This integrative CASP experiment was highlighted at the recent CASP13 meeting (http://www.predictioncenter.org/), where the carbohydrate-binding module (CBM56) of a β−1,3-glucanase from Bacillus circulans with 184 amino acids (18.9 kDa) was studied as the first FRET data-assisted target F0964. In CASP14, the single-model protein structure prediction by the artificial intelligence (AI) network AlphaFold2, which was developed by Google's AI offshoot DeepMind (https://deepmind.com), has approached perfection (Callaway, 2020). This deep-learning program combines the evolutionary information from multiple sequence alignments with structural information from the PDB for computing 3D structural models of a protein from its amino-acid sequence.

However, one has to be aware that many proteins do not only adopt their thermodynamically most stable conformation but frequently exist as ensembles of conformations that have high functional relevance. Thus, mapping dynamic ensembles represents the subsequent challenge of structural biology for the next decades. Due to their high time-resolution, smFRET-studies and integrative modeling can contribute a lot to solving this problem. We propose that members of the smFRET community who are interested in using smFRET for integrative structural biology participate in the CASP competition. Involvement could progress in several stages: (1) Predicting single- and multi-state structural models: the smFRET community will only submit distances that will be evaluated with respect to the known (but undisclosed) crystal structure(s). (2) Predicting ensembles as in the case of CBM56: for targets that are identified as difficult by the predictors and for which multiple possible folds are submitted without a clear winner, a FRET-assisted round could be insightful where the FRET distances distributions can be used as an experimental ‘ground truth’ for checking whether multiple conformations in an exchange are present.

These recommendations apply mostly to present and future practitioners of smFRET-driven integrative modeling. That being said, smFRET is one of many biophysical techniques that can provide experimental restraints in integrative modeling (XL-MS, single-particle cryoEM, NMR, SAXS). Therefore, we propose that, at a later stage, an all-biophysics integrative structural biology competition be established.

SmFRET meetings

Several gatherings of FRET practitioners at the Annual Biophysical Society Meetings, supported by the Biological Fluorescence subgroup, provided a platform for planning future activities and establishing the FRET community. As further joint actions, satellite meetings to the Conference on Methods and Applications in Fluorescence (MAF) have been organized to discuss practices, standards, competitions, and joint publications. We envision an occasional dedicated meeting for the smFRET community, such as the Bunsen meetings on FRET held in 2011 and the international discussion meeting in 2016 at the Max Planck Institute for Biophysical Chemistry in Göttingen, Germany (http://fret.uni-duesseldorf.de/cms/home.html). However, to open these meetings to smFRET practitioners outside of Europe, we propose to rotate the venue among continents. We also suggest using the satellite meetings and workshops to disseminate information (details of accurate FRET measurements, common practices, standards and competitions) and to give newcomers the chance to interact with the experienced researchers in the field.

Inspired by the online seminars emerging in response to the COVID-19 pandemic, smFRET webinars and web conferences open to all should be pursued. They provide FRET researchers the unique opportunity to attend and socialize virtually and would be a forum for good scientific practice of open science for the FRET community.

Special issues in journals

To further stimulate newcomers to engage in advanced smFRET experiments, the FRET community could benefit by hosting special issues in journals dedicated to data analyses (e.g., Data in Brief, Methods in Molecular Biology, or Nature Protocols). Here, various laboratories can describe typical datasets or protocols for the methods they have developed. Also, journals disseminating methodologies and protocols from A-to-Z via video recordings could be useful. For example, there is a special issue focusing on FRET planned in the Journal of Visualized Experiments (https://www.jove.com/methods-collections/682).

Future of smFRET

With improved communication and dissemination within the FRET community and agreement on the standard information required for depositing FRET-based or integrative structural data, smFRET will be better positioned to impact the expanding field of dynamic structural biology. We expect integrated approaches such as combining smFRET with NMR (Aznauryan et al., 2016; Liu et al., 2018; Milles et al., 2015; Sottini et al., 2020; Tsytlonok et al., 2019; Voith von Voithenberg et al., 2016), EPR (Boura et al., 2011; Masliah et al., 2018; Peter et al., 2020; Peulen et al., 2020; Sanabria et al., 2020; Vöpel et al., 2014), cross-linking mass spectrometry (Calabrese et al., 2020; Liu et al., 2018; Tyagi and Lemke, 2015), hydrogen/deuterium exchange (Calabrese et al., 2020; Liu et al., 2018; Munro and Lee, 2018), and/or MD simulations (see section Structural modeling and below) will have a big impact in the future.

One example of a major area of interest that is profiting from these developments is the study of intrinsically disordered proteins using smFRET experiments (Gomes and Gradinaru, 2017; Gomes et al., 2020; LeBlanc et al., 2018; Lee et al., 2018; Metskas and Rhoades, 2020; Nasir et al., 2020; Schuler et al., 2016). The dynamic nature of these proteins and their interactions play major roles in numerous cellular processes, including the formation of membrane-less intracellular biomolecular condensates, a new paradigm that presents huge challenges for traditional tools of structural biology (Banani et al., 2017; Choi et al., 2020; van der Lee et al., 2014; Wright and Dyson, 2015). Many IDPs undergo large folding transitions in conjunction with binding to partners, while others remain disordered upon complex formation (Schuler et al., 2020; Wang and Wang, 2019; Wu et al., 2017). SmFRET studies of these systems began more than a decade ago and have tackled increasingly complex systems using more advanced methods, including three-color smFRET or complex labeling schemes (Borgia et al., 2018; Kim and Chung, 2020; Lee et al., 2018; Metskas and Rhoades, 2020; Milles et al., 2015; Nasir et al., 2019; Schuler et al., 2016; Yoo et al., 2020). A recent study that combined smFRET, NMR, and MD simulations to investigate the interaction of H1 with ProTα is highlighted in Figure 9 (Borgia et al., 2018).

Using smFRET to investigate the structure and dynamics of ultrahigh-affinity IDP complexes.

(A) SmFRET efficiency histograms for FRET between a donor label (Alexa488) attached at various positions to the linker histone H1 (shown in blue) with the IDP ProTα (shown in red) labeled at different positions with the acceptor fluorophore (Alexa594). (B) For structural calculations of the H1-ProTα complex, coarse-grained MD simulations were performed. From the MD simulations, an ensemble of structures was determined. Eleven examples of configurations are shown and projected onto the first three principle components (PC1, PC2, and PC3) of the inter-residue distance map. 2D projections of the full ensemble are shown in gray (axes are labeled in Å). (C) A comparison of the experimental FRET efficiencies (filled squares) and the FRET efficiencies estimated from simulated structures (open circles) shows good agreement between the measured and simulated values. Pictograms indicate the variations of dye positions studied. (Panels A, B, and C: Copyright 2018, Nature Publishing Group, a division of Macmillan Publishers Limited. All rights reserved. Reproduced from Borgia et al., 2018, with permission. Further reproduction of this panel would need permission from the copyright holder.)

© 2018, Macmillan Publishers Limited, part of Springer Nature. All rights reserved. Panels A-C were originally published as Figure 3i, 4c and 4a in Borgia et al., 2018. Further reproduction of this panel would need permission from the copyright holder

Studies of IDPs are even more challenging in heterogeneous environments such as phase-separated mesoscale structures (e.g., membrane-less organelles) or cells (Nasir et al., 2019). Here, the strengths of smFRET would be especially valuable, while the field, at the same time, will benefit from the methodological developments. The ultimate goal is to combine both the structural and dynamic information in order to reduce the ambiguity in the underlying structures of conformational states and to gain detailed information on kinetic pathways between the associated states.

Although we have focused mostly on kinetic studies, smFRET-based structure determination and structural dynamics in this paper, there are a myriad of other new exciting directions where smFRET will have future impact. A detailed description of the various possibilities is beyond the scope of this report, but it is worth mentioning a few of them below.

Epilogue

In this article, we have summarized current perspectives on the status of the smFRET field, limitations that still need to be overcome, and joint efforts towards the adoption of consistent methodologies and open-science practices. While this article encourages a discussion regarding optimal smFRET practices, it is important to remember that, as scientists, we should value independence of thought and creativity. Hence, our recommendations should be taken as constructive suggestions, and it is important to realize that many biological questions can be answered using multiple approaches. On the one hand, the reproducibility and reliability of smFRET measurements are currently limited by the variety of approaches taken to calculate the FRET efficiency and the resulting inter-dye distance. Combining years of experience from various experts in an open discussion can help us, as a community, to improve the methodology and overcome many of its challenges. On the other hand, it is important to be open to new ideas and approaches. Here is where open scientific practices can help the community to quickly exchange data and analysis approaches to test new ideas. Such a community effort is necessary to consolidate the role of smFRET as a useful tool in various fields and to jointly move the field forward. Our hope is that these efforts will benefit not only the smFRET community, but also the structural biology community and science in general.

Appendix 1

Simulation software

References

  1. Book
    1. Becker W
    (2019)
    The bh TCSPC Handbook, Eighth Edition
    Becker & Hickl GmbH.
  2. Book
    1. Blanco M
    2. Walter NG
    (2010) Analysis of complex single-molecule FRET time trajectories
    In: Walter E, editors. Methods in Enzymology. Academic Press. pp. 153–178.
    https://doi.org/10.1016/S0076-6879(10)72011-5
  3. Conference
    1. Cohen AE
    2. Moerner WE
    (2005) The anti-Brownian electrophoretic trap (ABEL trap): fabrication and software
    Proceedings of SPIE 5699, Imaging, Manipulation, and Analysis of Biomolecules and Cells: Fundamentals and Applications III. pp. 296–305.
    https://doi.org/10.1117/12.598689
  4. Conference
    1. Hadzic M
    2. Kowerko D
    3. Börner R
    4. Zelger-Paulus S
    5. Sigel RKO
    (2016) Detailed analysis of complex single molecule FRET data with the software MASH
    Proc. SPIE 9711, Imaging, Manipulation, and Analysis of Biomolecules, Cells, and Tissues IX, 971119.
    https://doi.org/10.1117/12.2211191
  5. Software
    1. Nettels D
    2. Schuler B
    (2020) Fretica
    University of Zurich.
  6. Book
    1. Salem C-B
    2. Ploetz E
    3. Lamb DC
    (2019) Chapter 2 - Probing dynamics in single molecules
    In: Johnson CK, editors. Developments in Physical & Theoretical Chemistry, Spectroscopy and Dynamics of Single Molecules. Elsevier. pp. 71–115.
    https://doi.org/10.1016/B978-0-12-816463-1.00002-X
  7. Book
    1. Selvin PR
    2. Ha T
    (2008)
    Single-Molecule Techniques
    Cold Spring Harbor Laboratory Press.

Article and author information

Author details

  1. Eitan Lerner

    Department of Biological Chemistry, The Alexander Silberman Institute of Life Sciences, and The Center for Nanoscience and Nanotechnology, Faculty of Mathematics & Science, The Edmond J. Safra Campus, The Hebrew University of Jerusalem, Jerusalem, Israel
    Contribution
    Wrote the initial draft of the manuscript, coordinated the input from all authors and discussed the revisions
    Contributed equally with
    Anders Barth and Jelle Hendrix
    For correspondence
    eitan.lerner@mail.huji.ac.il
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-3791-5277
  2. Anders Barth

    Lehrstuhl für Molekulare Physikalische Chemie, Heinrich-Heine-Universität, Düsseldorf, Germany
    Contribution
    Expanded the initial draft, contributed extensively to the writing and revising of the manuscript, and prepared figures
    Contributed equally with
    Eitan Lerner and Jelle Hendrix
    For correspondence
    a.barth@tudelft.nl
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0003-3671-3072
  3. Jelle Hendrix

    Dynamic Bioimaging Lab, Advanced Optical Microscopy Centre and Biomedical Research Institute (BIOMED), Hasselt University, Diepenbeek, Belgium
    Contribution
    Expanded the initial draft, contributed extensively to the writing and revising of the manuscript, and prepared figures
    Contributed equally with
    Eitan Lerner and Anders Barth
    For correspondence
    jelle.hendrix@uhasselt.be
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0001-5731-1297
  4. Benjamin Ambrose

    Department of Chemistry, University of Sheffield, Sheffield, United Kingdom
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  5. Victoria Birkedal

    Department of Chemistry and iNANO center, Aarhus University, Aarhus, Denmark
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  6. Scott C Blanchard

    Department of Structural Biology, St. Jude Children's Research Hospital, Memphis, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0003-2717-9365
  7. Richard Börner

    Laserinstitut HS Mittweida, University of Applied Science Mittweida, Mittweida, Germany
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0001-8407-6624
  8. Hoi Sung Chung

    Laboratory of Chemical Physics, National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health, Bethesda, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  9. Thorben Cordes

    Physical and Synthetic Biology, Faculty of Biology, Ludwig-Maximilians-Universität München, Planegg-Martinsried, Germany
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-8598-5499
  10. Timothy D Craggs

    Department of Chemistry, University of Sheffield, Sheffield, United Kingdom
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-7121-0609
  11. Ashok A Deniz

    Department of Integrative Structural and Computational Biology, The Scripps Research Institute, La Jolla, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  12. Jiajie Diao

    Department of Cancer Biology, University of Cincinnati School of Medicine, Cincinnati, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  13. Jingyi Fei

    Department of Biochemistry and Molecular Biology and The Institute for Biophysical Dynamics, University of Chicago, Chicago, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-9775-3820
  14. Ruben L Gonzalez

    Department of Chemistry, Columbia University, New York, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-1344-5581
  15. Irina V Gopich

    Laboratory of Chemical Physics, National Institute of Diabetes and Digestive and Kidney Diseases, National Institutes of Health, Bethesda, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  16. Taekjip Ha

    Department of Biophysics and Biophysical Chemistry, Department of Biomedical Engineering, Johns Hopkins University School of Medicine, Howard Hughes Medical Institute, Baltimore, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0003-2195-6258
  17. Christian A Hanke

    Lehrstuhl für Molekulare Physikalische Chemie, Heinrich-Heine-Universität, Düsseldorf, Germany
    Contribution
    prepared figures and contributed to writing the manuscript
    Competing interests
    No competing interests declared
  18. Gilad Haran

    Department of Chemical and Biological Physics, Weizmann Institute of Science, Rehovot, Israel
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0003-1837-9779
  19. Nikos S Hatzakis

    1. Department of Chemistry & Nanoscience Centre, University of Copenhagen, Copenhagen, Denmark
    2. Denmark Novo Nordisk Foundation Centre for Protein Research, Faculty of Health and Medical Sciences, University of Copenhagen, Copenhagen, Denmark
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  20. Sungchul Hohng

    Department of Physics and Astronomy, and Institute of Applied Physics, Seoul National University, Seoul, Republic of Korea
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  21. Seok-Cheol Hong

    Center for Molecular Spectroscopy and Dynamics, Institute for Basic Science and Department of Physics, Korea University, Seoul, Republic of Korea
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  22. Thorsten Hugel

    Institute of Physical Chemistry and Signalling Research Centres BIOSS and CIBSS, University of Freiburg, Freiburg, Germany
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0003-3292-4569
  23. Antonino Ingargiola

    Department of Chemistry and Biochemistry, and Department of Physiology, University of California, Los Angeles, Los Angeles, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-9348-1397
  24. Chirlmin Joo

    Department of BioNanoScience, Kavli Institute of Nanoscience, Delft University of Technology, Delft, Netherlands
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  25. Achillefs N Kapanidis

    Biological Physics Research Group, Clarendon Laboratory, Department of Physics, University of Oxford, Oxford, United Kingdom
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  26. Harold D Kim

    School of Physics, Georgia Institute of Technology, Atlanta, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  27. Ted Laurence

    Physical and Life Sciences Directorate, Lawrence Livermore National Laboratory, Livermore, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  28. Nam Ki Lee

    School of Chemistry, Seoul National University, Seoul, Republic of Korea
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-6597-555X
  29. Tae-Hee Lee

    Department of Chemistry, Pennsylvania State University, University Park, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  30. Edward A Lemke

    1. Departments of Biology and Chemistry, Johannes Gutenberg University, Mainz, Germany
    2. Institute of Molecular Biology (IMB), Mainz, Germany
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-0634-0503
  31. Emmanuel Margeat

    Centre de Biologie Structurale (CBS), CNRS, INSERM, Universitié de Montpellier, Montpellier, France
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  32. Jens Michaelis

    Institüt of Biophysics, Ulm University, Ulm, Germany
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  33. Xavier Michalet

    Department of Chemistry and Biochemistry, and Department of Physiology, University of California, Los Angeles, Los Angeles, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0001-6602-7693
  34. Sua Myong

    Department of Biophysics, Johns Hopkins University, Baltimore, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  35. Daniel Nettels

    Department of Biochemistry and Department of Physics, University of Zurich, Zurich, Switzerland
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  36. Thomas-Otavio Peulen

    Department of Bioengineering and Therapeutic Sciences, University of California, San Francisco, San Francisco, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  37. Evelyn Ploetz

    Physical Chemistry, Department of Chemistry, Center for Nanoscience (CeNS), Center for Integrated Protein Science Munich (CIPSM) and Nanosystems Initiative Munich (NIM), Ludwig-Maximilians-Universität, München, Germany
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0003-0922-875X
  38. Yair Razvag

    Department of Biological Chemistry, The Alexander Silberman Institute of Life Sciences, and The Center for Nanoscience and Nanotechnology, Faculty of Mathematics & Science, The Edmond J. Safra Campus, The Hebrew University of Jerusalem, Jerusalem, Israel
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  39. Nicole C Robb

    Warwick Medical School, University of Warwick, Coventry, United Kingdom
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  40. Benjamin Schuler

    Department of Biochemistry and Department of Physics, University of Zurich, Zurich, Switzerland
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-5970-4251
  41. Hamid Soleimaninejad

    Biological Optical Microscopy Platform (BOMP), University of Melbourne, Parkville, Australia
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  42. Chun Tang

    College of Chemistry and Molecular Engineering, PKU-Tsinghua Center for Life Sciences, Beijing National Laboratory for Molecular Sciences, Peking University, Beijing, China
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  43. Reza Vafabakhsh

    Department of Molecular Biosciences, Northwestern University, Evanston, United States
    Contribution
    Contributed to writing the manuscript
    Competing interests
    No competing interests declared
  44. Don C Lamb

    Physical Chemistry, Department of Chemistry, Center for Nanoscience (CeNS), Center for Integrated Protein Science Munich (CIPSM) and Nanosystems Initiative Munich (NIM), Ludwig-Maximilians-Universität, München, Germany
    Contribution
    Expanded the initial draft, contributed extensively to the writing and revising of the manuscript
    For correspondence
    d.lamb@lmu.de
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-0232-1903
  45. Claus AM Seidel

    Lehrstuhl für Molekulare Physikalische Chemie, Heinrich-Heine-Universität, Düsseldorf, Germany
    Contribution
    Expanded the initial draft, contributed extensively to the writing and revising of the manuscript, and prepared figures
    For correspondence
    cseidel@hhu.de
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-5171-149X
  46. Shimon Weiss

    1. Department of Chemistry and Biochemistry, and Department of Physiology, University of California, Los Angeles, Los Angeles, United States
    2. Department of Physiology, CaliforniaNanoSystems Institute, University of California, Los Angeles, Los Angeles, United States
    Contribution
    Wrote the initial draft of the manuscript, coordinated the input from all authors and discussed the revisions
    For correspondence
    sweiss@chem.ucla.edu
    Competing interests
    No competing interests declared
    ORCID icon "This ORCID iD identifies the author of this article:" 0000-0002-0720-5426

Funding

National Institutes of Health (GM130942)

  • Eitan Lerner
  • Shimon Weiss

National Institutes of Health (GM095904)

  • Xavier Michalet

National Institutes of Health (GM079238)

  • Scott C Blanchard

National Institutes of Health (GM098859)

  • Scott C Blanchard

National Institutes of Health (GM084288)

  • Ruben L Gonzalez

National Institutes of Health (GM137608)

  • Ruben L Gonzalez

National Institutes of Health (GM112882)

  • Harold D Kim

National Institutes of Health (GM123164)

  • Tae-Hee Lee

National Institutes of Health (GM130793)

  • Tae-Hee Lee

National Institutes of Health (GM140272)

  • Reza Vafabakhsh

National Institutes of Health (GM130375)

  • Ashok A Deniz

National Institutes of Health (128185)

  • Jingyi Fei

National Institute of Diabetes and Digestive and Kidney Diseases

  • Hoi Sung Chung
  • Irina V Gopich

National Science Foundation (1818147)

  • Shimon Weiss

National Science Foundation (1842951)

  • Shimon Weiss

National Science Foundation (2004016)

  • Ruben L Gonzalez

Human Frontier Science Program (RGP0061/2019)

  • Shimon Weiss

European Research Council (638536)

  • Thorben Cordes

European Research Council (860954)

  • Thorben Cordes

European Research Council (SMPFv2.0)

  • Edward A Lemke

European Research Council (671208)

  • Claus AM Seidel

European Research Council (819299)

  • Chirlmin Joo

Deutsche Forschungsgemeinschaft (GRK2062)

  • Thorben Cordes

Deutsche Forschungsgemeinschaft (SFB863)

  • Thorben Cordes

Deutsche Forschungsgemeinschaft (PL696/4-1)

  • Evelyn Ploetz

Deutsche Forschungsgemeinschaft (SPP2191 402723784)

  • Edward A Lemke

Deutsche Forschungsgemeinschaft (SFB 1129 240245660)

  • Edward A Lemke

Deutsche Forschungsgemeinschaft (SE 1195/21-1 (SPP 2191))

  • Claus AM Seidel

Deutsche Forschungsgemeinschaft (CIBSS - EXC-2189 - project ID 390939984)

  • Thorsten Hugel

Wellcome Trust (110164/Z/15/Z)

  • Achillefs N Kapanidis

Swiss National Science Foundation

  • Benjamin Schuler

Biotechnology and Biological Sciences Research Council (BB/S008896/1)

  • Achillefs N Kapanidis

Biotechnology and Biological Sciences Research Council (BB/T008032/1)

  • Timothy D Craggs

Royal Society (RGS\R2\180405)

  • Timothy D Craggs

Royal Society (DKR00620)

  • Nicole C Robb

Royal Society (RGF\R1\180054)

  • Nicole C Robb

Agence Nationale de la Recherche (ANR-17-CE09-0026-02)

  • Emmanuel Margeat

Agence Nationale de la Recherche (ANR-18-CE11-0004-02)

  • Emmanuel Margeat

Agence Nationale de la Recherche (ANR-19-CE44-0009-02)

  • Emmanuel Margeat

Israel Science Foundation (1250/19)

  • Gilad Haran

Israel Science Foundation (3565/20)

  • Eitan Lerner

National Research Foundation of Korea (2019R1A2C2090896)

  • Nam Ki Lee

National Research Foundation of Korea (NRF-2019R1A2C2005209)

  • Sungchul Hohng

National Research Foundation of Korea (NRF-2019R1A2C1089808)

  • Seok-Cheol Hong

Independent Fund Denmark (6110-00623B)

  • Viktoria Birkedal

National Key Research and Development Program of China (2018YFA0507700)

  • Chun Tang

UHasselt BOF fund (R-10789)

  • Jelle Hendrix

Deutsche Forschungsgemeinschaft (SFB1035)

  • Don C Lamb

Milner Fund

  • Eitan Lerner

KU Leuven Special Research Fund (C14/16/053)

  • Jelle Hendrix

Carlsbergfondet (CF16-0797)

  • Nikos S Hatzakis

Villum Fonden (18333)

  • Nikos S Hatzakis

Novo Nordisk (NNF14CC0001)

  • Nikos S Hatzakis

Ludwig-Maximilians-Universität München

  • Timothy D Craggs

Carolina Cancer Center of Nanotechnology Excellence

  • Timothy D Craggs

University of Melbourne

  • Hamid Soleimaninejad

Searle Scholars Program

  • Jingyi Fei

University of Zurich

  • Richard Börner

EPSRC

  • Benjamin Ambrose

The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

Acknowledgements

We wish to thank Niko Hildebrandt and Sonja Schmid for fruitful discussions. We thank Niels Vandenberk, KU Leuven for use of a figure from his PhD thesis, and Bianca Herman, Universität Freiburg, for providing material for Figure 6. RB wishes to thank Fabio D Steffen for insightful discussions. This position paper was supported by the National Institutes of Health (NIH, grant R01 GM130942 to SW, and to EL as a subaward; grant R01 GM095904 to XM; grants R01 GM079238 and 7R01GM098859-09 to SCB; grants R01 GM084288 and R01 GM137608 to RLG; grant R01 GM112882 to HDK; grants R01 GM123164 and R01 GM130793 to THL; grant R01 GM140272 to RV; grant R35 GM130375 to AAD; New Innovator Award 1DP2GM128185-01 to JF), the Intramural Research Program of the National Institute of Diabetes and Digestive and Kidney Diseases, NIH (to HSC and IVG), the National Science Foundation (NSF, grants MCB-1818147 and MCB-1842951 to SW; grant CHE-2004016 to RLG), the Human Frontier Science Program (HFSP, grant RGP0061/2019 to SW), the European Research Council (ERC; grant numbers 638536 and 860954 to TC; grant SMPFv2.0 to EAL; grant No. 671208 (hybridFRET) to CAMS; grant No. 819299 to CJ), the Deutsche Forschungsgemeinschaft (DFG, grants GRK2062 (C03) and SFB863 (A13) to TC; grant (PL696/4-1) to EP; grants SPP2191 402723784 and SFB 1129 240245660 to EAL; grant SE 1195/21–1 (SPP 2191) to CAMS; Germany's Excellence Strategy, CIBSS – EXC-2189 – project ID 390939984 to T Hugel), the Wellcome Trust (grant 110164/Z/15/Z to AK), the Swiss National Science Foundation (to BS), the UK Biotechnology and Biological Sciences Research Council (grant BB/S008896/1 to AK; grant BB/T008032/1 to TDC), the Royal Society (grant RGS\R2\180405 to TDC; grants DKR00620 and RGF\R1\180054 to NCR), the Agence Nationale de la Recherche (grants ANR-17-CE09-0026-02, ANR-18-CE11-0004-02 and ANR-19-CE44-0009-02 to EM), the Israel Science Foundation (ISF, grant 1250/19 to GH; grant 3565/20 within the KillCorona – Curbing Coronavirus Research Program to EL), the National Research Foundation of Korea (grant 2019R1A2C2090896 to NKL; grant NRF-2019R1A2C2005209 to SH, grant NRF-2019R1A2C1089808 to SCH.), the Independent Fund Denmark (grant 6110-00623B to VB), the National Key R and D Program from the Ministry of Science and Technology of China (grant 2018YFA0507700 to CT), the Milner Fund (to EL), the Hebrew University of Jerusalem (start-up funds to EL), the DFG SFB1035, project A11 and Ludwig Maximillian Universität (LMU, the LMUinnovativ program BioImaging Network, BIN and Center for NanoScience (CeNS)) to DCL, the Biological Optical Microscopy Platform (BOMP) at University of Melbourne (to HS), the Searle Scholars Program (to JF), the University of Applied Sciences Mittweida and the Alfred-Werner-Legat of the Department of Chemistry, University Zurich (financial support to RB), the EPSRC DTP Studentship (to BA), the UHasselt BOF fund (R-10789 to JH), KU Leuven Special Research Fund (C14/16/053 to JH) and the CASLMU fellowship (to TC, SW, EL). Carlsberg Foundation Distinguished Associate Professor program CF16-0797, the Villum Foundation Center of Excellence BIONEC (grant 18333), and Novo Nordisk Foundation (grant NNF14CC0001) to Novo Nordisk Foundation Center for Protein Research (for NSH). GH holds the Hilda Pomeraniec Memorial Professorial Chair. This work was performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory under Contract DE-AC52-07NA27344 (for TL).

Senior and Reviewing Editor

  1. Olga Boudker, Weill Cornell Medicine, United States

Version history

  1. Received: June 29, 2020
  2. Accepted: February 9, 2021
  3. Version of Record published: March 29, 2021 (version 1)
  4. Version of Record updated: March 31, 2021 (version 2)
  5. Version of Record updated: April 9, 2021 (version 3)

Copyright

This is an open-access article, free of all copyright, and may be freely reproduced, distributed, transmitted, modified, built upon, or otherwise used by anyone for any lawful purpose. The work is made available under the Creative Commons CC0 public domain dedication.

Metrics

  • 11,817
    Page views
  • 2,161
    Downloads
  • 123
    Citations

Article citation count generated by polling the highest count across the following sources: Crossref, Scopus, PubMed Central.

Download links

A two-part list of links to download the article, or parts of the article, in various formats.

Downloads (link to download the article as PDF)

Open citations (links to open the citations from this article in various online reference manager services)

Cite this article (links to download the citations from this article in formats compatible with various reference manager tools)

  1. Eitan Lerner
  2. Anders Barth
  3. Jelle Hendrix
  4. Benjamin Ambrose
  5. Victoria Birkedal
  6. Scott C Blanchard
  7. Richard Börner
  8. Hoi Sung Chung
  9. Thorben Cordes
  10. Timothy D Craggs
  11. Ashok A Deniz
  12. Jiajie Diao
  13. Jingyi Fei
  14. Ruben L Gonzalez
  15. Irina V Gopich
  16. Taekjip Ha
  17. Christian A Hanke
  18. Gilad Haran
  19. Nikos S Hatzakis
  20. Sungchul Hohng
  21. Seok-Cheol Hong
  22. Thorsten Hugel
  23. Antonino Ingargiola
  24. Chirlmin Joo
  25. Achillefs N Kapanidis
  26. Harold D Kim
  27. Ted Laurence
  28. Nam Ki Lee
  29. Tae-Hee Lee
  30. Edward A Lemke
  31. Emmanuel Margeat
  32. Jens Michaelis
  33. Xavier Michalet
  34. Sua Myong
  35. Daniel Nettels
  36. Thomas-Otavio Peulen
  37. Evelyn Ploetz
  38. Yair Razvag
  39. Nicole C Robb
  40. Benjamin Schuler
  41. Hamid Soleimaninejad
  42. Chun Tang
  43. Reza Vafabakhsh
  44. Don C Lamb
  45. Claus AM Seidel
  46. Shimon Weiss
(2021)
FRET-based dynamic structural biology: Challenges, perspectives and an appeal for open-science practices
eLife 10:e60416.
https://doi.org/10.7554/eLife.60416

Share this article

https://doi.org/10.7554/eLife.60416

Further reading

    1. Biochemistry and Chemical Biology
    2. Developmental Biology
    Zhi Li, Yuedi Wang ... Zeyang Zhou
    Research Article

    Imidacloprid is a global health threat that severely poisons the economically and ecologically important honeybee pollinator, Apis mellifera. However, its effects on developing bee larvae remain largely unexplored. Our pilot study showed that imidacloprid causes developmental delay in bee larvae, but the underlying toxicological mechanisms remain incompletely understood. In this study, we exposed bee larvae to imidacloprid at environmentally relevant concentrations of 0.7, 1.2, 3.1, and 377 ppb. There was a marked dose-dependent delay in larval development, characterized by reductions in body mass, width, and growth index. However, imidacloprid did not affect on larval survival and food consumption. The primary toxicological effects induced by elevated concentrations of imidacloprid (377 ppb) included inhibition of neural transmission gene expression, induction of oxidative stress, gut structural damage, and apoptosis, inhibition of developmental regulatory hormones and genes, suppression of gene expression levels involved in proteolysis, amino acid transport, protein synthesis, carbohydrate catabolism, oxidative phosphorylation, and glycolysis energy production. In addition, we found that the larvae may use antioxidant defenses and P450 detoxification mechanisms to mitigate the effects of imidacloprid. Ultimately, this study provides the first evidence that environmentally exposed imidacloprid can affect the growth and development of bee larvae by disrupting molting regulation and limiting the metabolism and utilization of dietary nutrients and energy. These findings have broader implications for studies assessing pesticide hazards in other juvenile animals.

    1. Biochemistry and Chemical Biology
    2. Cancer Biology
    Pengfei Guo, Rebecca C Lim ... Hui Zhang
    Research Article Updated

    The Polycomb Repressive Complex 2 (PRC2) methylates H3K27 to regulate development and cell fate by transcriptional silencing. Alteration of PRC2 is associated with various cancers. Here, we show that mouse Kdm1a deletion causes a dramatic reduction of PRC2 proteins, whereas mouse null mutation of L3mbtl3 or Dcaf5 results in PRC2 accumulation and increased H3K27 trimethylation. The catalytic subunit of PRC2, EZH2, is methylated at lysine 20 (K20), promoting EZH2 proteolysis by L3MBTL3 and the CLR4DCAF5 ubiquitin ligase. KDM1A (LSD1) demethylates the methylated K20 to stabilize EZH2. K20 methylation is inhibited by AKT-mediated phosphorylation of serine 21 in EZH2. Mouse Ezh2K20R/K20R mutants develop hepatosplenomegaly associated with high GFI1B expression, and Ezh2K20R/K20R mutant bone marrows expand hematopoietic stem cells and downstream hematopoietic populations. Our studies reveal that EZH2 is regulated by methylation-dependent proteolysis, which is negatively controlled by AKT-mediated S21 phosphorylation to establish a methylation-phosphorylation switch to regulate the PRC2 activity and hematopoiesis.