Minflux

Last updated

MINFLUX, or minimal fluorescence photon fluxes microscopy, is a super-resolution light microscopy method that images and tracks objects in two and three dimensions with single-digit nanometer resolution. [1] [2] [3]

Contents

MINFLUX uses a structured excitation beam with at least one intensity minimum – typically a doughnut-shaped beam with a central intensity zero – to elicit photon emission from a fluorophore. The position of the excitation beam is controlled with sub-nanometer precision, and when the intensity zero is positioned exactly on the fluorophore, the system records no emission. Thus, the system requires few emitted photons to determine the fluorophore's location with high precision. In practice, overlapping the intensity zero and the fluorophore would require a priori location knowledge to position the beam. As this is not the case, the excitation beam is moved around in a defined pattern to probe the emission from the fluorophore near the intensity minimum. [1]

Each localization takes less than 5 microseconds, [1] so MINFLUX can construct images of nanometric structures or track single molecules in fixed and live specimens by pooling the locations of fluorescent labels. Because the goal is to locate the point where a fluorophore stops emitting, MINFLUX significantly reduces the number of fluorescence photons needed for localization compared to other methods. [2] [4]

Principle

MINFLUX overcomes the Abbe diffraction limit in light microscopy and distinguishes individual fluorescing molecules by leveraging the photophysical properties of fluorophores. The system temporarily silences (sets in an OFF-state) all but one molecule within a diffraction-limited area (DLA) and then locates that single active (in an ON-state) molecule. [1] Super-resolution microscopy techniques like stochastic optical reconstruction microscopy (STORM) and photoactivated localization microscopy (PALM) do the same. [5] However, MINFLUX differs in how it determines the molecule’s location.

The excitation beam used in MINFLUX has a local intensity minimum or intensity zero. The position of this intensity zero in a sample is adjusted via control electronics and actuators with sub-nanometer spatial and sub-microsecond temporal precision. When the active molecule located at is in a non-zero intensity area of the excitation beam, it fluoresces. The number of photons emitted by the active molecule is proportional to the excitation beam intensity at that position.

When an active fluorescent molecule is in the vicinity of the intensity zero of the MINFLUX excitation beam, the number of photons it emits is proportional to the intensity of the excitation beam at that position. The intensity of that emission can be approximated by a quadratic function. In this one-dimensional diagram, recording photon emission with the excitation beam at two different positions enables estimating the intensity minimum between the two points, corresponding to the active molecule's estimated position. MINFLUX principle (described in one dimension).png
When an active fluorescent molecule is in the vicinity of the intensity zero of the MINFLUX excitation beam, the number of photons it emits is proportional to the intensity of the excitation beam at that position. The intensity of that emission can be approximated by a quadratic function. In this one-dimensional diagram, recording photon emission with the excitation beam at two different positions enables estimating the intensity minimum between the two points, corresponding to the active molecule's estimated position.

In the vicinity of the excitation beam intensity zero, the intensity of the emission from the active molecule when the intensity zero is located at position can be approximated by a quadratic function. Therefore, the recorded number of emission photons is:

where  is a measure of the collection efficiency of detection, the absorption cross-section of the emitter, and the quantum yield of fluorescence.

In other words, photon fluxes emitted by the active molecule when it is located close to the zero-intensity point of the excitation beam carry information about its distance to the center of the beam. That information can be used to find the position of the active molecule. The position is probed with a set of excitation intensities . For example, the active molecule is excited with the same doughnut-shaped beam moved to different positions. The probing results in a corresponding set of photon counts . These photon counts are probabilistic; each time such a set is measured, the result is a different realization of photon numbers fluctuating around a mean value. Since their distribution follows Poissonian statistics, the expected position of the active molecule can be estimated from the photon numbers, using, for example, a maximum likelihood estimation of the form:

The position  maximizes the likelihood that the measured set of photon counts occurred exactly as recorded and is thus, an estimate of the active molecule’s location. [6]

Localization process

MINFLUX localizes an active fluorescent molecule using a probing scheme. The doughnut-shaped excitation beam is positioned at different points at the periphery and in the middle of a probing area L. The active molecule is excited at each position, and photon fluxes are recorded. MINFLUX uses the flux patterns of these recordings to reposition the excitation beam to center on the active molecule and then performs another probing iteration. With each iteration, the probing area L is constricted, and the intensity zero of the excitation beam more accurately overlaps with the position of the active molecule. Localization of a fluorophore using MINFLUX.png
MINFLUX localizes an active fluorescent molecule using a probing scheme. The doughnut-shaped excitation beam is positioned at different points at the periphery and in the middle of a probing area L. The active molecule is excited at each position, and photon fluxes are recorded. MINFLUX uses the flux patterns of these recordings to reposition the excitation beam to center on the active molecule and then performs another probing iteration. With each iteration, the probing area L is constricted, and the intensity zero of the excitation beam more accurately overlaps with the position of the active molecule.

Recordings of the emitting active molecule at two different excitation beam positions are needed to use the quadratic approximation in the one-dimensional basic principle described above. Each recording provides a one-dimensional distance value to the center of the excitation beam. In two dimensions, at least three recording points are needed to ascertain a location that can be used to move the MINFLUX excitation beam toward the target molecule. These recording points demarcate a probing area L. Balzarotti et al. [1] use the Cramér-Rao limit to show that constricting this probing area significantly improves localization precision, more so than increasing the number of emitted photons:

where is the Cramér-Rao limit, is the diameter of the probing area, and is the number of emitted photons.

MINFLUX takes advantage of this feature when localizing an active fluorophore. It records photon fluxes using a probing scheme of at least three recording points around the probing area and one point at the center. These fluxes differ at each recording point as the active molecule is excited by different light intensities. Those flux patterns inform the repositioning of the probing area to center on the active molecule. Then the probing process is repeated. With each probing iteration, MINFLUX constricts the probing area , narrowing the space where the active molecule can be located. Thus, the distance remaining between the intensity zero and the active molecule is determined more precisely at each iteration. The steadily improving positional information minimizes the number of fluorescence photons and the time that MINFLUX needs to achieve precise localizations. [7]

Applications

By pooling the determined locations of multiple fluorescent molecules in a specimen, MINFLUX generates images of nanoscopic structures with a resolution of 1–3 nm. [8] MINFLUX has been used to image DNA origami [1] [9] and the nuclear pore complex [10] and to elucidate the architecture of subcellular structures in mitochondria and photoreceptors. [11] [12] Because MINFLUX does not collect large numbers of photons emitted from target molecules, localization is faster than with conventional camera-based systems. [13] Thus, MINFLUX can iteratively localize the same molecule at microsecond intervals over a defined period. MINFLUX has been used to track the movement of the motor protein kinesin-1, both in vitro and in vivo, [14] [15] and to monitor configurational changes of the mechanosensitive ion channel PIEZO1. [16]

See also

Related Research Articles

<span class="mw-page-title-main">Microscopy</span> Viewing of objects which are too small to be seen with the naked eye

Microscopy is the technical field of using microscopes to view objects and areas of objects that cannot be seen with the naked eye. There are three well-known branches of microscopy: optical, electron, and scanning probe microscopy, along with the emerging field of X-ray microscopy.

<span class="mw-page-title-main">Fluorophore</span> Agents that emit light after excitation by light

A fluorophore is a fluorescent chemical compound that can re-emit light upon light excitation. Fluorophores typically contain several combined aromatic groups, or planar or cyclic molecules with several π bonds.

A total internal reflection fluorescence microscope (TIRFM) is a type of microscope with which a thin region of a specimen, usually less than 200 nanometers can be observed.

<span class="mw-page-title-main">Fluorescence microscope</span> Optical microscope that uses fluorescence and phosphorescence

A fluorescence microscope is an optical microscope that uses fluorescence instead of, or in addition to, scattering, reflection, and attenuation or absorption, to study the properties of organic or inorganic substances. "Fluorescence microscope" refers to any microscope that uses fluorescence to generate an image, whether it is a simple set up like an epifluorescence microscope or a more complicated design such as a confocal microscope, which uses optical sectioning to get better resolution of the fluorescence image.

Fluorescence-lifetime imaging microscopy or FLIM is an imaging technique based on the differences in the exponential decay rate of the photon emission of a fluorophore from a sample. It can be used as an imaging technique in confocal microscopy, two-photon excitation microscopy, and multiphoton tomography.

<span class="mw-page-title-main">Two-photon excitation microscopy</span> Fluorescence imaging technique

Two-photon excitation microscopy is a fluorescence imaging technique that is particularly well-suited to image scattering living tissue of up to about one millimeter in thickness. Unlike traditional fluorescence microscopy, where the excitation wavelength is shorter than the emission wavelength, two-photon excitation requires simultaneous excitation by two photons with longer wavelength than the emitted light. The laser is focused onto a specific location in the tissue and scanned across the sample to sequentially produce the image. Due to the non-linearity of two-photon excitation, mainly fluorophores in the micrometer-sized focus of the laser beam are excited, which results in the spatial resolution of the image. This contrasts with confocal microscopy, where the spatial resolution is produced by the interaction of excitation focus and the confined detection with a pinhole.

Fluorescence correlation spectroscopy (FCS) is a statistical analysis, via time correlation, of stationary fluctuations of the fluorescence intensity. Its theoretical underpinning originated from L. Onsager's regression hypothesis. The analysis provides kinetic parameters of the physical processes underlying the fluctuations. One of the interesting applications of this is an analysis of the concentration fluctuations of fluorescent particles (molecules) in solution. In this application, the fluorescence emitted from a very tiny space in solution containing a small number of fluorescent particles (molecules) is observed. The fluorescence intensity is fluctuating due to Brownian motion of the particles. In other words, the number of the particles in the sub-space defined by the optical system is randomly changing around the average number. The analysis gives the average number of fluorescent particles and average diffusion time, when the particle is passing through the space. Eventually, both the concentration and size of the particle (molecule) are determined. Both parameters are important in biochemical research, biophysics, and chemistry.

<span class="mw-page-title-main">STED microscopy</span> Technique in fluorescence microscopy

Stimulated emission depletion (STED) microscopy is one of the techniques that make up super-resolution microscopy. It creates super-resolution images by the selective deactivation of fluorophores, minimizing the area of illumination at the focal point, and thus enhancing the achievable resolution for a given system. It was developed by Stefan W. Hell and Jan Wichmann in 1994, and was first experimentally demonstrated by Hell and Thomas Klar in 1999. Hell was awarded the Nobel Prize in Chemistry in 2014 for its development. In 1986, V.A. Okhonin had patented the STED idea. This patent was unknown to Hell and Wichmann in 1994.

RESOLFT, an acronym for REversible Saturable OpticaLFluorescence Transitions, denotes a group of optical fluorescence microscopy techniques with very high resolution. Using standard far field visible light optics a resolution far below the diffraction limit down to molecular scales can be obtained.

Fluorescence interference contrast (FLIC) microscopy is a microscopic technique developed to achieve z-resolution on the nanometer scale.

<span class="mw-page-title-main">Vertico spatially modulated illumination</span>

Vertico spatially modulated illumination (Vertico-SMI) is the fastest light microscope for the 3D analysis of complete cells in the nanometer range. It is based on two technologies developed in 1996, SMI and SPDM. The effective optical resolution of this optical nanoscope has reached the vicinity of 5 nm in 2D and 40 nm in 3D, greatly surpassing the λ/2 resolution limit applying to standard microscopy using transmission or reflection of natural light according to the Abbe resolution limit That limit had been determined by Ernst Abbe in 1873 and governs the achievable resolution limit of microscopes using conventional techniques.

<span class="mw-page-title-main">Fluorescence in the life sciences</span> Scientific investigative technique

Fluorescence is used in the life sciences generally as a non-destructive way of tracking or analysing biological molecules. Some proteins or small molecules in cells are naturally fluorescent, which is called intrinsic fluorescence or autofluorescence. Alternatively, specific or general proteins, nucleic acids, lipids or small molecules can be "labelled" with an extrinsic fluorophore, a fluorescent dye which can be a small molecule, protein or quantum dot. Several techniques exist to exploit additional properties of fluorophores, such as fluorescence resonance energy transfer, where the energy is passed non-radiatively to a particular neighbouring dye, allowing proximity or protein activation to be detected; another is the change in properties, such as intensity, of certain dyes depending on their environment allowing their use in structural studies.

Super-resolution microscopy is a series of techniques in optical microscopy that allow such images to have resolutions higher than those imposed by the diffraction limit, which is due to the diffraction of light. Super-resolution imaging techniques rely on the near-field or on the far-field. Among techniques that rely on the latter are those that improve the resolution only modestly beyond the diffraction-limit, such as confocal microscopy with closed pinhole or aided by computational methods such as deconvolution or detector-based pixel reassignment, the 4Pi microscope, and structured-illumination microscopy technologies such as SIM and SMI.

Photo-activated localization microscopy and stochastic optical reconstruction microscopy (STORM) are widefield fluorescence microscopy imaging methods that allow obtaining images with a resolution beyond the diffraction limit. The methods were proposed in 2006 in the wake of a general emergence of optical super-resolution microscopy methods, and were featured as Methods of the Year for 2008 by the Nature Methods journal. The development of PALM as a targeted biophysical imaging method was largely prompted by the discovery of new species and the engineering of mutants of fluorescent proteins displaying a controllable photochromism, such as photo-activatible GFP. However, the concomitant development of STORM, sharing the same fundamental principle, originally made use of paired cyanine dyes. One molecule of the pair, when excited near its absorption maximum, serves to reactivate the other molecule to the fluorescent state.

Super-resolution optical fluctuation imaging (SOFI) is a post-processing method for the calculation of super-resolved images from recorded image time series that is based on the temporal correlations of independently fluctuating fluorescent emitters.

Super-resolution dipole orientation mapping (SDOM) is a form of fluorescence polarization microscopy (FPM) that achieved super resolution through polarization demodulation. It was first described by Karl Zhanghao and others in 2016. Fluorescence polarization (FP) is related to the dipole orientation of chromophores, making fluorescence polarization microscopy possible to reveal structures and functions of tagged cellular organelles and biological macromolecules. In addition to fluorescence intensity, wavelength, and lifetime, the fourth dimension of fluorescence—polarization—can also provide intensity modulation without the restriction to specific fluorophores; its investigation in super-resolution microscopy is still in its infancy.

Three-photon microscopy (3PEF) is a high-resolution fluorescence microscopy based on nonlinear excitation effect. Different from two-photon excitation microscopy, it uses three exciting photons. It typically uses 1300 nm or longer wavelength lasers to excite the fluorescent dyes with three simultaneously absorbed photons. The fluorescent dyes then emit one photon whose energy is three times the energy of each incident photon. Compared to two-photon microscopy, three-photon microscopy reduces the fluorescence away from the focal plane by , which is much faster than that of two-photon microscopy by . In addition, three-photon microscopy employs near-infrared light with less tissue scattering effect. This causes three-photon microscopy to have higher resolution than conventional microscopy.

<span class="mw-page-title-main">Fluorescence imaging</span> Type of non-invasive imaging technique

Fluorescence imaging is a type of non-invasive imaging technique that can help visualize biological processes taking place in a living organism. Images can be produced from a variety of methods including: microscopy, imaging probes, and spectroscopy.

<span class="mw-page-title-main">Non-degenerate two-photon absorption</span> Simultaneous absorption of two photons of differing energies by a molecule

In atomic physics, non-degenerate two-photon absorption or two-color two-photon excitation is a type of two-photon absorption (TPA) where two photons with different energies are (almost) simultaneously absorbed by a molecule, promoting a molecular electronic transition from a lower energy state to a higher energy state. The sum of the energies of the two photons is equal to, or larger than, the total energy of the transition.

Francisco Balzarotti is an Argentinian scientist known for his work in super-resolution microscopy, particularly MINFLUX. He is a Group Leader at the Research Institute of Molecular Pathology (IMP) in Vienna, Austria.

References

  1. 1 2 3 4 5 6 Balzarotti, Francisco; Eilers, Yvan; Gwosch, Klaus C.; Gynnå, Arvid H.; Westphal, Volker; Stefani, Fernando D.; Elf, Johan; Hell, Stefan W. (2017-02-10). "Nanometer resolution imaging and tracking of fluorescent molecules with minimal photon fluxes". Science. 355 (6325): 606–612. arXiv: 1611.03401 . Bibcode:2017Sci...355..606B. doi:10.1126/science.aak9913. hdl:11858/00-001M-0000-002C-2D9A-4. ISSN   0036-8075. PMID   28008086. S2CID   5418707.
  2. 1 2 Sigal, Yaron M.; Zhou, Ruobo; Zhuang, Xiaowei (2018-08-31). "Visualizing and discovering cellular structures with super-resolution microscopy". Science. 361 (6405): 880–887. Bibcode:2018Sci...361..880S. doi:10.1126/science.aau1044. ISSN   0036-8075. PMC   6535400 . PMID   30166485.
  3. Tang, Verena (2023-01-10). "Ein Potenzial, das noch in keiner Form gehoben ist". Spektrum.de. Retrieved 2023-11-17.
  4. Xiao, Jie; Ha, Taekjip (2017-02-10). "Flipping nanoscopy on its head". Science. 355 (6325): 582–584. Bibcode:2017Sci...355..582X. doi:10.1126/science.aam5409. ISSN   0036-8075. PMC   8989063 . PMID   28183938.
  5. Lelek, Mickaël; Gyparaki, Melina T.; Beliu, Gerti; Schueder, Florian; Griffié, Juliette; Manley, Suliana; Jungmann, Ralf; Sauer, Markus; Lakadamyali, Melike; Zimmer, Christophe (2021-06-03). "Single-molecule localization microscopy". Nature Reviews Methods Primers. 1 (1). doi:10.1038/s43586-021-00038-x. ISSN   2662-8449. PMC   9160414 . PMID   35663461.
  6. Eilers, Y. 2017. Localizing and tracking of fluorescent molecules with minimal photon fluxes. Doctoral Dissertation. Georg-August-University Göttingen. https://ediss.uni-goettingen.de/handle/11858/00-1735-0000-002E-E329-2 (accessed January 2024)
  7. Gwosch, Klaus C.; Pape, Jasmin K.; Balzarotti, Francisco; Hoess, Philipp; Ellenberg, Jan; Ries, Jonas; Hell, Stefan W. (2020-02-02). "MINFLUX nanoscopy delivers 3D multicolor nanometer resolution in cells". Nature Methods. 17 (2): 217–224. doi:10.1038/s41592-019-0688-0. ISSN   1548-7091. PMID   31932776. S2CID   210168754.
  8. Bond, Charles; Santiago-Ruiz, Adriana N.; Tang, Qing; Lakadamyali, Melike (2022-01-20). "Technological advances in super-resolution microscopy to study cellular processes". Molecular Cell. 82 (2): 315–332. doi:10.1016/j.molcel.2021.12.022. PMC   8852216 . PMID   35063099.
  9. Wogan, Tim (2017-01-09). "New super-resolution microscope combines Nobel-winning technologies". Physics World. Retrieved 2023-11-17.
  10. Schmidt, Roman; Weihs, Tobias; Wurm, Christian A.; Jansen, Isabelle; Rehman, Jasmin; Sahl, Steffen J.; Hell, Stefan W. (2021-03-05). "MINFLUX nanometer-scale 3D imaging and microsecond-range tracking on a common fluorescence microscope". Nature Communications. 12 (1): 1478. Bibcode:2021NatCo..12.1478S. doi:10.1038/s41467-021-21652-z. ISSN   2041-1723. PMC   7935904 . PMID   33674570.
  11. Pape, Jasmin K.; Stephan, Till; Balzarotti, Francisco; Büchner, Rebecca; Lange, Felix; Riedel, Dietmar; Jakobs, Stefan; Hell, Stefan W. (2020-08-25). "Multicolor 3D MINFLUX nanoscopy of mitochondrial MICOS proteins". Proceedings of the National Academy of Sciences. 117 (34): 20607–20614. Bibcode:2020PNAS..11720607P. doi: 10.1073/pnas.2009364117 . ISSN   0027-8424. PMC   7456099 . PMID   32788360.
  12. Grabner, Chad P.; Jansen, Isabelle; Neef, Jakob; Weihs, Tobias; Schmidt, Roman; Riedel, Dietmar; Wurm, Christian A.; Moser, Tobias (2022-07-15). "Resolving the molecular architecture of the photoreceptor active zone with 3D-MINFLUX". Science Advances. 8 (28): eabl7560. Bibcode:2022SciA....8L7560G. doi:10.1126/sciadv.abl7560. ISSN   2375-2548. PMC   9286502 . PMID   35857490.
  13. Strack, Rita (2017-02-28). "Taking nanoscopy to the limit". Nature Methods. 14 (3): 221. doi:10.1038/nmeth.4211. ISSN   1548-7091.
  14. Deguchi, Takahiro; Iwanski, Malina K.; Schentarra, Eva-Maria; Heidebrecht, Christopher; Schmidt, Lisa; Heck, Jennifer; Weihs, Tobias; Schnorrenberg, Sebastian; Hoess, Philipp; Liu, Sheng; Chevyreva, Veronika; Noh, Kyung-Min; Kapitein, Lukas C.; Ries, Jonas (2023-03-10). "Direct observation of motor protein stepping in living cells using MINFLUX". Science. 379 (6636): 1010–1015. Bibcode:2023Sci...379.1010D. doi:10.1126/science.ade2676. ISSN   0036-8075. PMC   7614483 . PMID   36893247.
  15. Dambeck, Susanne (2017-03-23). "New Super Tool for Cell Biology". Lindau Nobel Laureate Meetings. Retrieved 2024-01-22.
  16. Mulhall, Eric M.; Gharpure, Anant; Lee, Rachel M.; Dubin, Adrienne E.; Aaron, Jesse S.; Marshall, Kara L.; Spencer, Kathryn R.; Reiche, Michael A.; Henderson, Scott C.; Chew, Teng-Leong; Patapoutian, Ardem (2023-08-31). "Direct observation of the conformational states of PIEZO1". Nature. 620 (7976): 1117–1125. Bibcode:2023Natur.620.1117M. doi:10.1038/s41586-023-06427-4. ISSN   0028-0836. PMC   10468401 . PMID   37587339.