Open Access
8 March 2024 Molecular mapping of neuronal architecture using STORM microscopy and new fluorescent probes for SMLM imaging
Victor Breton, Paul Nazac, David Boulet, Lydia Danglot
Author Affiliations +
Abstract

Imaging neuronal architecture has been a recurrent challenge over the years, and the localization of synaptic proteins is a frequent challenge in neuroscience. To quantitatively detect and analyze the structure of synapses, we recently developed free SODA software to detect the association of pre and postsynaptic proteins. To fully take advantage of spatial distribution analysis in complex cells, such as neurons, we also selected some new dyes for plasma membrane labeling. Using Icy SODA plugin, we could detect and analyze synaptic association in both conventional and single molecule localization microscopy, giving access to a molecular map at the nanoscale level. To replace those molecular distributions within the neuronal three-dimensional (3D) shape, we used MemBright probes and 3D STORM analysis to decipher the entire 3D shape of various dendritic spine types at the single-molecule resolution level. We report here the example of synaptic proteins within neuronal mask, but these tools have a broader spectrum of interest since they can be used whatever the proteins or the cellular type. Altogether with SODA plugin, MemBright probes thus provide the perfect toolkit to decipher a nanometric molecular map of proteins within a 3D cellular context.

1.

Introduction

Imaging neuronal architecture has been a recurrent challenge over the years. The use of Golgi technique by Ramón y Cajal paved the way for the first characterization of neuronal architecture using microscopy on fixed brains. Indeed, metallic impregnation with silver salts provided an opportunity to see and reconstruct the dendritic architecture of various types of neurons in the depth of the nervous tissue. Although Golgi staining is still used in widefield microscopy, its use in confocal microscopy remains limited for 3D analysis. The use of fluorescent labeling in conjunction with 3D microscopy led to the production of large amounts of published data that are on the way to being classified and accessible through various infrastructures or free repositories (eBrains, Zenodo, etc.). This huge amount of data and their accessibility raise the question of potential new automated, unbiased statistical analysis.

2.

Colocalization and Coupling Analysis in Conventional and Super-Resolution Microscopy

To analyze the structure of synapses quantitatively, we recently developed, in association with statisticians, free software to detect the association of pre- and postsynaptic proteins. This software, called SODA for standard object distance analysis, makes it possible to identify and measure the spatial distribution of either clusters (conventional microscopy) or single molecules (single molecule localization microscopy) and provides the distance of association when those are statically found associated.1,2 SODA (available here; see Ref. 3) can be used in conventional microscopy (confocal, widefield, and video microscopy) or in super-resolution microscopy, such as SIM, STED, or even SMLM (PALM or STORM). From the mathematical point of view, SODA will analyze the cellular shape and spot density (Fig. 1) to evaluate the expected spatial distribution using Ripley’s function. If clusters are more frequently associated than the expected random distribution, then they are identified as associated spots. As a proof of concept, we analyzed the distribution of three synaptic molecules named synapsin, homer, and PSD-95 using SIM microscopy. Using only 15 pictures, we were able to analyze about 50,000 synapses and identify that the distance between a synapsin and post-synaptic PSD95 cluster was 107±73  nm while the PSD-95-homer was 64±48  nm. Beyond raw distances, this system allows to detect in a quantitative manner any morphological variations that may occur in various mutants, physiological conditions or in certain synaptopathies.

Fig. 1

Workflow of the Easy SODA protocol available in Icy Software. Colocalization or distant association can be analyzed using the user-friendly Easy SODA protocol that is freely available in Icy Software. This protocol is a graphical programming automatization routine that allows analysis of synaptic proteins’ distribution within neuronal cell shape. Here, neurons are labeled with two proteins (green and cyan channels) that are distributed in clusters. Clusters are segmented using wavelet segmentation through “spot detector” plugin. Cell shape is extracted using a MAP2 stain, and segmentation is done using the “Hierarchical KMeans” plugin. Cluster distribution within the cellular mask is then analyzed through the “SODA” plugin using Ripley’s analysis. Statistical associations are detected if any, and the proportion of associated clusters with their distance is provided with a p-value indicating the statistical robustness of the association. If many pictures are analyzed in batch mode, all results can be exported to Excel files. A molecular map is exported for each picture with an association color code. For example, if we take a red-green spot analysis: isolated green spots remain green, green spots associated with red are cyan, isolated red spots remain red, and red spots associated with green are pink. Localization of significant associations is thus visible at a glance over the cell mask (here in deep blue).

NPh_11_1_014414_f001.png

SODA can be used either with sparse labeling (>30100 objects per image)4 or with high-density labeling as in single-molecule localization microscopy1,2 where several thousands of localizations can be retrieved. The only limitation of SODA is the need to get the cell boundary to correctly evaluate the object’s density. In contrast to methods using Voronoï tesselation5 that are limited up to now two-dimensional analysis, SODA can be used in 3D, which is an added value to analyze thick 3D volume STORM images. Because SODA does not rely on any overlap methods, it is far less sensitive to high-density false positive colocalization artifacts.1,2

SODA can also be used for all other associations (either direct or distant) in neurosciences610 and even outside this field like, for example, in cell biology,1118 virology,12 in bacteria19 or plants.20 Its use has been highlighted in several reviews.2022

3.

Imaging Plasma Membrane in Live or Fixed Cells

3.1.

Optimizing Live Plasma Membrane Imaging with MemBright Probes

In order to be able to place the molecules within the cell shape, and in collaboration with chemists, we have selected new membrane probes capable of revealing the cell shape and imaging fine structures, such as dendritic spines.23

These membrane probes, named MemBright, upon insertion in the membrane, emit fluorescence with narrow emission peaks, allowing correlation with other conventional fluorescence for multi-color labeling. A family of seven members is now available and can be used all over the fluorescence spectrum (from 480 and 750 nm). Upon incubation with living cells, these probes insert directly into the membrane through a lipid anchor and thus reveal the cell shapes without the use of any transfection or viral vectors (Fig. 2). This strategy thus makes it possible to reveal all neurons and/or glial cells in a few minutes without any toxicity.

Fig. 2

Confocal image (20×) of hippocampal neurons after 2 months in a culture labeled with Cy3.5-MemBright probe reveals dendritic and axonal branch complexity. Scale bar: 100  μm.

NPh_11_1_014414_f002.png

The big advantage of MemBright probes is also its fluorogenic property. Indeed, MemBright probes are non-fluorescent within media and become fluorescent when reaching the plasma membrane. It means that the probe can be let within the cell culture media without having a fluorescence background under the microscope. Letting the probe in the cell chamber during live acquisition allows the perpetual replacement of the probes if bleaching occurs, thus leading to persistent bright labeling of the membrane.

MemBright probes are more efficient when incubated on cells in the absence of serum to avoid any titration of the probes by serum fat. In neurons and glial cells, we usually incubate MemBright probes in Krebs Ringer solution at a concentration of 200 nM at 37 deg under the microscope. The absence of serum optimizes the labeling, and the absence of phenol red lowers the fluorescence background. Fluorescence on the plasma membrane appears very fast within a few minutes. To avoid saturation of the plasma membrane with a huge amount of lipids, it is crucial not to use a high concentration of probes. MemBright probes are sufficiently bright to be used at the nanomolar range, whereas other commercial probes have to be used at the micromolar range. Moreover, it should be stressed that illumination of any fluorescent probes may induce the production of reactive oxygen species that can imbalance the intracellular redox state and be deleterious to the cell.24 Thus, it is also a good practice to minimize illumination time and frequency, to the minimal amount needed to the right sampling of a biological event. Previously, we could follow neuronal growth over time during 13 h, imaging every 2 min with low laser power (0.2% of an 561 laser line of an Elira PS1), without any detrimental effects.23 At last, we have shown that MemBright probes are resistant to permeabilization when fixed properly with a mixture of 4% paraformaldehyde-0.2%glutaraldehyde, and can thus be combined with conventional immunolabeling with primary and secondary antibodies.23 MemBright staining can be correlated either to live antibody staining or with immunochemistry on fixed samples. We could show double labeling of the plasma membrane and live L1-CAM endocytosed antibodies. We also showed that intracellular vesicular transporter VGLUT could be revealed by immunochemistry within 3D neuronal cell shape reconstructed using MemBright. This property thus allows identifying surface or internal protein locations using MemBright counterstaining to visualize cell shape. We have selected MemBright probes for their stability at the plasma membrane and their slow endocytosis. However, on long-term incubation, they will be finally endocytosed and can thus be used to label endocytic pathways [Fig. 3(c)].25

Fig. 3

Confocal images (93×) of various cell types labeled with Cy3.5-MemBright probe. (a) 3D rendering of E.coli bacteria incubated overnight with MemBright probes. (b) 3D rendering of HeLa cells incubated 30 min with MemBright. (c) Confocal section of hippocampal neurons incubated several hours with MemBright, revealing intracellular vesicles.

NPh_11_1_014414_f003.png

3.2.

Imaging Plasma Membrane of Various Cell Types

We originally showed that MemBright probes could be used in various cell types, such as epithelial cells in culture (HeLa cells or KB cells),23,2628 and dissociated hippocampal neurons,23 hippocampal astrocytes.23 We could also use MemBright probes to label live brain (hippocampus, cortex, and cerebellum) or liver slices, allowing the labeling in depth and imaging using confocal or two photons microscopy.23 Since our first paper in 2019, MemBright probes have been used by several other labs and cited in more than 60 articles and 39 reviews.2957 It has been used in B lymphocytes,58 in A431 cells,59 and reused in neuronal cells to label growth cone and initial segments of hippocampal neurons,60 presynaptic terminals,29 and post-synaptic compartments.53 It has also been used to label apoptotic bodies (AB), microvesicles (MV), and small EV (sEV) isolated from MIN6 pancreatic beta cells exposed to inflammatory, hypoxic, or genotoxic stressors.61 Since we were asked several times if MemBright could be used on bacteria, we did recently E. coli live labeling with MemBright – CY3.5. As shown in Fig. 3(a), the bacteria 3D shape can be efficiently labeled in live.

3.3.

Imaging Extracellular Vesicles

MemBright has been widely adopted by the extracellular vesicles community51 to track extracellular vesicles both in vitro or in vivo6181 in hippocampal63 or cortical neurons,79 zebrafish,62,67,72 breast cancer cells or tumours,66,71,80 myotubes,82 and red blood cells.74,76

Particle size distribution and zeta potential analysis of EVs derived from A375 cells using nanoparticle tracking analysis (NTA) showed that EVs labeled before and after labeling by MemBright have almost no change in size and only a slight shift of zeta potential.75 Due to its ease of use and brightness, MemBright has thus been widely used to label exosomes. However, it should be stressed that MemBright is not specific to extracellular vesicle labeling. MemBright will be able to label any membrane in contact with the probe. That means that membrane debris trails left behind by migrating cells will be labeled, whatever the nature of the membrane (EVs or not). Any membranous organelles (tubules, endosomes, lysosomes, synaptosomes, etc.) that would have been retrieved by ultracentrifugation can be labeled when incubated with MemBright. Some controls are thus needed before labeling to ensure that the fraction is homogeneous and not contaminated by different organelles.

Hyenne et al.62 show that MemBright can be used in pulse-chase experiments and that some CD63-GFP EVs can be labeled with MemBright, while others are not.62 Sung et al. concluded that MemBright can label exosomes as well as plasma membrane-derived EVs, but that MemBright does not label all exosomes.47 Using a pulse-chase experiment, it is expected that not all endosomal compartments will be labeled. Indeed, only those deriving from the plasma membrane exposed to the MemBright at a time t will be visible. It is likely that a proportion of EVs that were generated before or after incubation with MemBright and that are then stored in the cell will not be labeled by the MemBright wave. It is therefore essential to properly calibrate the labeling time incubation and the chase time to observe the desired events.

4.

Imaging Plasma Membrane with Single Molecule Localization Microscopy

MemBright probes can be used in conventional and super-resolution microscopy (Fig. 4 and Video 1) and thus make it possible to observe the molecular distribution of synaptic proteins in correlation with the structural morphology at the nanoscale using STORM. Altogether with the SODA plugin, MemBright probes thus provide the perfect tools to access a nanometric molecular map of various proteins within the 3D cellular context. These high-resolution techniques were set up on fixed samples, but it is expected that the need for super-resolution imaging with small molecular probes to allow imaging of living samples will be growing and will stimulate the development of new molecular probes. Indeed, chemical development will be required in the next years, since molecular probes are still up to now far beyond the recent development in instrumentation that can reach a resolution of 3 to 4 nm.83 One interesting track will be probably the development of self-blinking dyes8486 or the development of new convertible fluorophores28 for live SMLM. We are now working on a photoswitchable version of the MemBright that would be able to be photoswitched without any reducing buffer, to be used in live single molecule localization microscopy.87 Moreover since MemBright has been used a lot in the extracellular vesicle community, it indicates that membrane dyes are of great interest for people working on vesicular trafficking. One other challenge, in the next years, would be to develop new MemBright probes devoted to fast internalization, to be able to decipher different vesicular pathways with various colors.

Fig. 4

STORM imaging of hippocampal neurons labeled with Cy3.5-MemBright probe and imaged in conventional widefield microscopy [fire LUT in (a)] and in STORM microscopy (blue spheres in rectangle). (b) Magnification of the plasma membrane 3D STORM image shows the single molecule organization of MemBright all over the plasma membrane. Light blue localizations are closer and deep blue are deeper. (c)–(e) Example of stubby and mushroom dendritic spines in 3D STORM. All the localizations found in (c) (published previously in a different form in Ref. 3) can be used to reconstruct the 3D shape of the spine in a wireframe. This 3D shape can then be used for volumetric estimation or fine measurements of the spine neck.

NPh_11_1_014414_f004.png

5.

Appendix: Supplementary Information

Video 1 Conventional and STORM imaging of hippocampal neurons labeled with Cy3.5-MemBright to label dendritic spines in 3D at the nanoscale level (MP4, 60.9 MB [URL: https://doi.org/10.1117/1.NPh.11.1.014414.s1]).

Disclosures

The authors have no competing interests to declare and are not financially interested or remunerated for MemBright probe sales.

Code and Data Availability

Tutorial and download concerning SODA plugin and user-friendly “Easy SODA protocol,” can be found on Icy Software website here: https://icy.bioimageanalysis.org/protocol/easy-soda-2-colors-1-image/

Sample image is also available here on Zenodo: https://zenodo.org/record/4323312-.Y_3tqx1CdHQ

A French Tutorial on Ripley’s function and Icy SODA plugin can be found on YouTube: https://www.youtube.com/watch?v=7yVp73s-4TA

Acknowledgments

We would like to thank the NeurImag Core Facility team (part of IPNP, Inserm U. 1266 and Université Paris Cité) and member of the National Infrastructure France-BioImaging supported by the French National Research Agency (ANR-10-INBS-04) for their technical and scientific support. We thank the Leducq Foundation for supporting the acquisition of the Leica SP8 Confocal/STED 3DX microscope and DIM Elicit and FRC/Rotary for funding the Bruker Vutara STORM and Optera spinning disc systems, respectively. VB, DB, and PN were supported by the FLAG-ERA (Grant Senseï by ANR-19-HBPR-0003), ANR-19-CE16-0012-03, and ANR-22-CE17-010-02. This work was supported by grants from the Agence Nationale de la Recherche (ANR-19-CE16-0012-03; ANR-22-CE17-0010-02; ANR-21-CE11-0010-03) and FLAG-ERA Grant Senseï by ANR-19-HBPR-0003 to LD.

References

1. 

T. Lagache et al., “Mapping molecular assemblies with fluorescence microscopy and object-based spatial statistics,” Nat. Commun., 9 (1), 698 https://doi.org/10.1038/s41467-018-03053-x (2018). Google Scholar

2. 

T. Lagache et al., “Statistical analysis of molecule colocalization in bioimaging,” Cytometry A, 87 (6), 568 –579 https://doi.org/10.1002/cyto.a.22629 (2015). Google Scholar

4. 

T. Lagache, V. Meas-Yedid and J. C. Olivo-Marin, “A statistical analysis of spatial colocalization using Ripley’s K function,” in IEEE 10th Int. Symp. Biomed. Imaging, 896 –901 (2013). https://doi.org/10.1109/ISBI.2013.6556620 Google Scholar

5. 

F. Levet et al., “SR-Tesseler: a method to segment and quantify localization-based super-resolution microscopy data,” Nat. Methods, 12 (11), 1065 –1071 https://doi.org/10.1038/nmeth.3579 (2015). Google Scholar

6. 

Y. Badawi and H. Nishimune, “Super-resolution microscopy for analyzing neuromuscular junctions and synapses,” Neurosci. Lett., 715 134644 https://doi.org/10.1016/j.neulet.2019.134644 (2020). Google Scholar

7. 

A. Bodzeta, F. Berger and H. D. MacGillavry, “Subsynaptic mobility of presynaptic mGluR types is differentially regulated by intra- and extracellular interactions,” Mol. Biol. Cell, 33 (8), ar66 https://doi.org/10.1091/mbc.E21-10-0484 (2022). Google Scholar

8. 

G. Gagliano et al., “Light sheet illumination for 3D single-molecule super-resolution imaging of neuronal synapses,” Front. Synaptic Neurosci., 13 761530 https://doi.org/10.3389/fnsyn.2021.761530 (2021). Google Scholar

9. 

A. Gallo et al., “Role of the Sec22b-E-Syt complex in neurite growth and ramification,” J. Cell Sci., 133 (18), jcs247148 https://doi.org/10.1242/jcs.247148 (2020). Google Scholar

10. 

S. Lenoir et al., “Pridopidine rescues BDNF/TrkB trafficking dynamics and synapse homeostasis in a Huntington disease brain-on-a-chip model,” Neurobiol. Dis., 173 105857 https://doi.org/10.1016/j.nbd.2022.105857 (2022). Google Scholar

11. 

T. Burrinha et al., “Upregulation of APP endocytosis by neuronal aging drives amyloid-dependent synapse loss,” J. Cell Sci., 134 (9), jcs255752 https://doi.org/10.1242/jcs.255752 (2021). Google Scholar

12. 

G. Cosentino et al., “Respiratory syncytial virus ribonucleoproteins hijack microtubule Rab11 dependent transport for intracellular trafficking,” PLoS Pathogens, 18 (7), e1010619 https://doi.org/10.1371/journal.ppat.1010619 (2022). Google Scholar

13. 

M. L. Ferrari et al., “Shigella promotes major alteration of gut epithelial physiology and tissue invasion by shutting off host intracellular transport,” Proc. Natl. Acad. Sci. U. S. A., 116 (27), 13582 –13591 https://doi.org/10.1073/pnas.1902922116 (2019). Google Scholar

14. 

P. Cristofari et al., “Nanoscopic distribution of VAChT and VGLUT3 in striatal cholinergic varicosities suggests colocalization and segregation of the two transporters in synaptic vesicles,” Front. Mol. Neurosci., 15 991732 https://doi.org/10.3389/fnmol.2022.991732 (2022). Google Scholar

15. 

A. Grassart et al., “Bioengineered human organ-on-chip reveals intestinal microenvironment and mechanical forces impacting Shigella infection,” Cell Host Microbe, 26 (3), 435 –444.e4 https://doi.org/10.1016/j.chom.2019.08.007 (2019). Google Scholar

16. 

B. Gu et al., “Opposing effects of cohesin and transcription on CTCF organization revealed by super-resolution imaging,” Mol. Cell, 80 (4), 699 –711.e7 https://doi.org/10.1016/j.molcel.2020.10.001 (2020). Google Scholar

17. 

R. Kamiyama et al., “Cell-type-specific, multicolor labeling of endogenous proteins with split fluorescent protein tags in Drosophila,” Proc. Natl. Acad. Sci. U. S. A., 118 (23), e2024690118 https://doi.org/10.1073/pnas.2024690118 (2021). Google Scholar

18. 

G. F. W. Walpole et al., “Inactivation of Rho GTPases by Burkholderia cenocepacia induces a WASH-mediated actin polymerization that delays phagosome maturation,” Cell Rep., 31 (9), 107721 https://doi.org/10.1016/j.celrep.2020.107721 (2020). Google Scholar

19. 

I. Altinoglu, C. J. Merrifield and Y. Yamaichi, “Single molecule super-resolution imaging of bacterial cell pole proteins with high-throughput quantitative analysis pipeline,” Sci. Rep., 9 6680 https://doi.org/10.1038/s41598-019-43051-7 (2019). Google Scholar

20. 

L. Colin et al., “Imaging the living plant cell: from probes to quantification,” Plant Cell, 34 (1), 247 –272 https://doi.org/10.1093/plcell/koab237 (2022). Google Scholar

21. 

A. O. Khan and J. A. Pike, “Super-resolution imaging and quantification of megakaryocytes and platelets,” Platelets, 31 (5), 559 –569 https://doi.org/10.1080/09537104.2020.1732321 (2020). Google Scholar

22. 

A. Mancebo et al., “Efficient cross-correlation filtering of one- and two-color single molecule localization microscopy data,” Front. Bioinf., 1 739769 https://doi.org/10.3389/fbinf.2021.739769 (2021). Google Scholar

23. 

M. Collot et al., “MemBright: a family of fluorescent membrane probes for advanced cellular imaging and neuroscience,” Cell Chem. Biol., 26 (4), 600 –614.e7 https://doi.org/10.1016/j.chembiol.2019.01.009 (2019). Google Scholar

24. 

M. Marchetti et al., “Genetically encoded biosensors for the fluorescence detection of O2 and reactive O2 species,” Sensors, 23 (20), 8517 https://doi.org/10.3390/s23208517 (2023). Google Scholar

25. 

M. Collot et al., “Probing polarity and heterogeneity of lipid droplets in live cells using a push-pull fluorophore,” Anal. Chem., 91 (3), 1928 –1935 https://doi.org/10.1021/acs.analchem.8b04218 (2019). Google Scholar

26. 

S. Michelis et al., “Imaging and measuring vesicular acidification with a plasma membrane-targeted ratiometric pH probe,” Anal. Chem., 94 (15), 5996 –6003 https://doi.org/10.1021/acs.analchem.2c00574 (2022). Google Scholar

27. 

M. Collot et al., “Ultrabright and fluorogenic probes for multicolor imaging and tracking of lipid droplets in cells and tissues,” J. Am. Chem. Soc., 140 (16), 5401 –5411 https://doi.org/10.1021/jacs.7b12817 (2018). Google Scholar

28. 

L. Saladin et al., “Dual-color photoconvertible fluorescent probes based on directed photooxidation induced conversion for bioimaging,” Angew. Chem. Int. Ed. Engl., 62 (4), e202215085 https://doi.org/10.1002/anie.202215085 (2023). Google Scholar

29. 

C. B. Harper and K. J. Smillie, “Current molecular approaches to investigate pre-synaptic dysfunction,” J. Neurochem., 157 (2), 107 –129 https://doi.org/10.1111/jnc.15316 (2021). Google Scholar

30. 

K. K. Hu et al., “Chemical analysis of single cells and organelles,” Anal. Chem., 93 (1), 41 –71 https://doi.org/10.1021/acs.analchem.0c04361 (2021). Google Scholar

31. 

H. R. Jia et al., “Cell surface-localized imaging and sensing,” Chem. Soc. Rev., 50 (10), 6240 –6277 https://doi.org/10.1039/D1CS00067E (2021). Google Scholar

32. 

K. Kikuchi et al., “Photochemical mechanisms of fluorophores employed in single-molecule localization microscopy,” Angew. Chem. Int. Edit., 62 (1), 25 https://doi.org/10.1002/anie.202204745 (2023). Google Scholar

33. 

H. J. Kim and S. H. Jo, “Nighttime administration of antihypertensive medication: a review of chronotherapy in hypertension,” Kor. J. Intern. Med., 10 10 https://doi.org/10.3904/kjim.2023.304 (2023). Google Scholar

34. 

A. S. Klymchenko, “Fluorescent probes for lipid membranes: from the cell surface to organelles,” Accounts Chem. Res., 56 (1), 1 –12 https://doi.org/10.1021/acs.accounts.2c00586 (2023). Google Scholar

35. 

S. A. A. Kooijmans, O. G. de Jong and R. M. Schiffelers, “Exploring interactions between extracellular vesicles and cells for innovative drug delivery system design,” Adv. Drug Deliv. Rev., 173 252 –278 https://doi.org/10.1016/j.addr.2021.03.017 (2021). Google Scholar

36. 

R. Kundu, A. Chandra and A. Datta, “Fluorescent chemical tools for tracking anionic phospholipids,” Isr. J. Chem., 61 (3–4), 199 –216 https://doi.org/10.1002/ijch.202100003 (2021). Google Scholar

37. 

C. L. Liu et al., “Advances in the development of fluorescence probes for cell plasma membrane imaging,” Trends Anal. Chem., 133 116092 https://doi.org/10.1016/j.trac.2020.116092 (2020). Google Scholar

38. 

C. Magliaro et al., “Gotta Trace ’em All: a mini-review on tools and procedures for segmenting single neurons toward deciphering the structural connectome,” Front. Bioeng. Biotechnol., 7 (8), 202 https://doi.org/10.3389/fbioe.2019.00202 (2019). Google Scholar

39. 

S. Mansuri, T. Mukherjee and S. Kanvah, “Fluorescent sterol probes for intracellular transport, imaging, and therapeutics,” Curr. Opin. Chem. Biol., 71 (8), 102222 https://doi.org/10.1016/j.cbpa.2022.102222 (2022). Google Scholar

40. 

A. Méndez-Ardoy, J. J. Reina and J. Montenegro, “Synthesis and supramolecular functional assemblies of ratiometric pH probes,” Chem. Eur. J., 26 (34), 7516 –7536 https://doi.org/10.1002/chem.201904834 (2020). Google Scholar

41. 

S. Mukherjee et al., “Imaging viral infection by fluorescence microscopy: focus on HIV-1 early stage,” Viruses, 13 (2), 213 https://doi.org/10.3390/v13020213 (2021). Google Scholar

42. 

V. G. Pivovarenko and A. S. Klymchenko, “Fluorescent probes based on charge and proton transfer for probing biomolecular environment,” Chem. Rec., 30 e202300321 https://doi.org/10.1002/tcr.202300321 (2023). Google Scholar

43. 

C. Salomon et al., “Extracellular vesicles and their emerging roles as cellular messengers in endocrinology: an endocrine society scientific statement,” Endocr. Rev., 43 (3), 441 –468 https://doi.org/10.1210/endrev/bnac009 (2022). Google Scholar

44. 

S. Samanta et al., “Xanthene, cyanine, oxazine and BODIPY: the four pillars of the fluorophore empire for super-resolution bioimaging,” Chem. Soc. Rev., 52 (20), 7197 –7261 https://doi.org/10.1039/D2CS00905F (2023). Google Scholar

45. 

F. Schneider, H. Colin-York and M. Fritzsche, “Quantitative bio-imaging tools to dissect the interplay of membrane and cytoskeletal actin dynamics in immune cells,” Front. Immunol., 11 (13), 612542 https://doi.org/10.3389/fimmu.2020.612542 (2021). Google Scholar

46. 

Y. Sun et al., “Single-molecule detection-based super-resolution imaging in single-cell analysis: Inspiring progress and future prospects,” Trends Anal. Chem., 167 (18), 117255 https://doi.org/10.1016/j.trac.2023.117255 (2023). Google Scholar

47. 

B. H. Sung, C. A. Parent and A. M. Weaver, “Extracellular vesicles: critical players during cell migration,” Dev. Cell, 56 (13), 1861 –1874 https://doi.org/10.1016/j.devcel.2021.03.020 (2021). Google Scholar

48. 

T. Sych et al., “How does liquid-liquid phase separation in model membranes reflect cell membrane heterogeneity?,” Membranes, 11 (5), 323 https://doi.org/10.3390/membranes11050323 (2021). Google Scholar

49. 

K. L. Tosheva et al., “Between life and death: strategies to reduce phototoxicity in super-resolution microscopy,” J. Phys. D-Appl. Phys., 53 (16), 163001 https://doi.org/10.1088/1361-6463/ab6b95 (2020). Google Scholar

50. 

E. E. Ural et al., “Visualizing extracellular vesicles and their function in 3D tumor microenvironment models,” Int. J. Mol. Sci., 22 (9), 4784 https://doi.org/10.3390/ijms22094784 (2021). Google Scholar

51. 

F. J. Verweij et al., “The power of imaging to understand extracellular vesicle biology in vivo,” Nat. Methods, 18 (9), 1013 –1026 https://doi.org/10.1038/s41592-021-01206-3 (2021). Google Scholar

52. 

J. A. Welsh et al., “A compendium of single extracellular vesicle flow cytometry,” J. Extracell. Vesicles, 12 (2), 70 https://doi.org/10.1002/jev2.12299 (2023). Google Scholar

53. 

C. Werner, M. Sauer and C. Geis, “Super-resolving microscopy in neuroscience,” Chem. Rev., 121 (19), 11971 –12015 https://doi.org/10.1021/acs.chemrev.0c01174 (2021). Google Scholar

54. 

Y. L. Xing et al., “Analysis of extracellular vesicles as emerging theranostic nanoplatforms,” Coord. Chem. Rev., 424 213506 https://doi.org/10.1016/j.ccr.2020.213506 (2020). Google Scholar

55. 

S. Xu et al., “Molecular engineering of near-infrared fluorescent probes for cell membrane imaging,” Molecules, 28 (4), 1906 https://doi.org/10.3390/molecules28041906 (2023). Google Scholar

56. 

L. Yang et al., “Small-molecule fluorescent probes for plasma membrane staining: design, mechanisms and biological applications,” Coord. Chem. Rev., 474 214862 https://doi.org/10.1016/j.ccr.2022.214862 (2023). Google Scholar

57. 

C. Zhou, G. W. N. Chia and K. T. Yong, “Membrane-intercalating conjugated oligoelectrolytes,” Chem. Soc. Rev., 51 (24), 9917 –9932 https://doi.org/10.1039/D2CS00014H (2022). Google Scholar

58. 

H. Y. Kwon et al., “Lipid-oriented live-cell distinction of B and T lymphocytes,” J. Am. Chem. Soc., 143 (15), 5836 –5844 https://doi.org/10.1021/jacs.1c00944 (2021). Google Scholar

59. 

C. Charpentier et al., “Ultrabright terbium nanoparticles for FRET biosensing and in situ imaging of epidermal growth factor receptors,” Chem. Eur. J., 26 (64), 14602 –14611 https://doi.org/10.1002/chem.202002007 (2020). Google Scholar

60. 

A. Honda et al., “Very-long-chain fatty acids are crucial to neuronal polarity by providing sphingolipids to lipid rafts,” Cell Rep., 42 (10), 113195 https://doi.org/10.1016/j.celrep.2023.113195 (2023). Google Scholar

61. 

K. R. Giri et al., “Molecular and functional diversity of distinct subpopulations of the stressed insulin-secreting cell’s vesiculome,” Front. Immunol., 11 1814 https://doi.org/10.3389/fimmu.2020.01814 (2020). Google Scholar

62. 

V. Hyenne et al., “Studying the fate of tumor extracellular vesicles at high spatiotemporal resolution using the zebrafish embryo,” Dev. Cell, 48 (4), 554 –572.e7 https://doi.org/10.1016/j.devcel.2019.01.014 (2019). Google Scholar

63. 

A. Antoniou et al., “Neuronal extracellular vesicles and associated microRNAs induce circuit connectivity downstream BDNF,” Cell Rep., 42 (2), 112063 https://doi.org/10.1016/j.celrep.2023.112063 (2023). Google Scholar

64. 

M. J. Boyer et al., “Endothelial cell-derived extracellular vesicles alter vascular smooth muscle cell phenotype through high-mobility group box proteins,” J. Extracell. Vesicles, 9 (1), 19 https://doi.org/10.1080/20013078.2020.1781427 (2020). Google Scholar

65. 

M. Dehghani et al., “Systematic evaluation of PKH labelling on extracellular vesicle size by nanoparticle tracking analysis,” Sci. Rep., 10 (1), 10 https://doi.org/10.1038/s41598-020-66434-7 (2020). Google Scholar

66. 

E. Gargiulo et al., “Extracellular vesicle secretion by leukemia cells in vivo promotes CLL progression by hampering antitumor t-cell responses,” Blood Cancer Discov., 4 (1), 54 –77 https://doi.org/10.1158/2643-3230.BCD-22-0029 (2023). Google Scholar

67. 

S. Ghoroghi et al., “Ral GTPases promote breast cancer metastasis by controlling biogenesis and organ targeting of exosomes,” Elife, 10 e61539 https://doi.org/10.7554/eLife.61539 (2021). Google Scholar

68. 

Y. He et al., “Fluorescence labeling of extracellular vesicles for diverse bio-applications in vitro and in vivo,” Chem. Commun., 59 (44), 6609 –6626 https://doi.org/10.1039/D3CC00998J (2023). Google Scholar

69. 

S. R. Jung et al., “Error-correction method for high-throughput sizing of nanoscale vesicles with single-molecule localization microscopy,” J. Phys. Chem. B, 127 (12), 2701 –2707 https://doi.org/10.1021/acs.jpcb.2c09053 (2023). Google Scholar

70. 

R. Lees et al., “Single extracellular vesicle transmembrane protein characterization by nano-flow cytometry,” J. Vis. Exp., (185), 15 https://doi.org/10.3791/64020 (2022). Google Scholar

71. 

L. Loconte et al., “Detection of the interactions of tumour derived extracellular vesicles with immune cells is dependent on EV-labelling methods,” J. Extracell. Vesicles, 12 (12), 15 https://doi.org/10.1002/jev2.12384 (2023). Google Scholar

72. 

B. Mary et al., “Blood flow diverts extracellular vesicles from endothelial degradative compartments to promote angiogenesis,” EMBO Rep., 24 e57042 https://doi.org/10.15252/embr.202357042 (2023). Google Scholar

73. 

G. E. Melling et al., “Confocal microscopy analysis reveals that only a small proportion of extracellular vesicles are successfully labelled with commonly utilised staining methods,” Sci. Rep., 12 (1), 13 https://doi.org/10.1038/s41598-021-04225-4 (2022). Google Scholar

74. 

A. Musicò et al., “Surface functionalization of extracellular vesicle nanoparticles with antibodies: a first study on the protein corona “variable”,” Nanoscale Adv., 5 (18), 4703 –4717 https://doi.org/10.1039/D3NA00280B (2023). Google Scholar

75. 

Y. A. Ren et al., “Rapid enrichment and sensitive detection of extracellular vesicles through measuring the phospholipids and transmembrane protein in a microfluidic chip,” Biosens. Bioelectron., 199 113870 https://doi.org/10.1016/j.bios.2021.113870 (2022). Google Scholar

76. 

B. V. Rodriguez et al., “An ex vivo model of interactions between extracellular vesicles and peripheral mononuclear blood cells in whole blood,” J. Extracell. Vesicles, 12 (12), 12368 https://doi.org/10.1002/jev2.12368 (2023). Google Scholar

77. 

Y. Sato, K. Ohira and S. Nishizawa, “Self-assembly and disassembly of membrane curvature-sensing peptide-based deep-red fluorescent probe for highly sensitive sensing of exosomes,” ACS Sens., 5 522 –526 https://doi.org/10.1021/acssensors.2c02498 (2023). Google Scholar

78. 

T. Shimomura et al., “New lipophilic fluorescent dyes for labeling extracellular vesicles: characterization and monitoring of cellular uptake,” Bioconjugate Chem., 32 (4), 680 –684 https://doi.org/10.1021/acs.bioconjchem.1c00068 (2021). Google Scholar

79. 

J. Solana-Balaguer et al., “Neuron-derived extracellular vesicles contain synaptic proteins, promote spine formation, activate TrkB-mediated signalling and preserve neuronal complexity,” J. Extracell. Vesicles, 12 (9), e12355 https://doi.org/10.1002/jev2.12355 (2023). Google Scholar

80. 

M. Tkach et al., “Extracellular vesicles from triple negative breast cancer promote pro-inflammatory macrophages associated with better clinical outcome,” Proc. Natl. Acad. Sci. U. S. A., 119 (17), e2107394119 https://doi.org/10.1073/pnas.2107394119 (2022). Google Scholar

81. 

J. Y. Zhu and G. C. Bazan, “Molecular orientation and optimization of membrane dyes based on conjugated oligoelectrolytes,” Cell Rep. Phys. Sci., 4 (6), 101429 https://doi.org/10.1016/j.xcrp.2023.101429 (2023). Google Scholar

82. 

E. Lemerle et al., “Caveolae and Bin1 form ring-shaped platforms for T-tubule initiation,” Elife, 12 e84139 https://doi.org/10.7554/eLife.84139 (2023). Google Scholar

83. 

R. Schmidt et al., “MINFLUX nanometer-scale 3D imaging and microsecond-range tracking on a common fluorescence microscope,” Nat. Commun., 12 (1), 1478 https://doi.org/10.1038/s41467-021-21652-z (2021). Google Scholar

84. 

M. Remmel et al., “Accelerated MINFLUX nanoscopy, through spontaneously fast-blinking fluorophores,” Small, 19 (12), e2206026 https://doi.org/10.1002/smll.202206026 (2023). Google Scholar

85. 

M. Remmel et al., “Spontaneously blinking fluorophores for accelerated MINFLUX nanoscopy,” (2022). Google Scholar

86. 

K. Grussmayer et al., “Self-blinking dyes unlock high-order and multiplane super-resolution optical fluctuation imaging,” ACS Nano, 14 (7), 9156 –9165 https://doi.org/10.1021/acsnano.0c04602 (2020). Google Scholar

87. 

L. Saladin et al., “Targeted photoconvertible BODIPYs based on directed photooxidation induced conversion for applications in photoconversion and live super resolution imaging,” (2023). Google Scholar

Biography

Victor Breton has been a PhD student in cell biology at the Institut de Psychiatrie et Neurosciences de Paris since September 2019. He graduated in physiology and neurosciences from the University of Paris-Saclay. His PhD concerns membrane nanostructures, which require the use of microscopes exceeding the light diffraction limit. To tackle those questions, he worked on membrane probes compatible with nanoscopy and set up specific image analysis protocols.

Paul Nazac started his PhD in cell biology and neuroscience at the Institut de Psychiatrie et Neurosciences de Paris in October 2022. He graduated in cell biology from Sorbonne Université and also graduated in cell signaling and integrative neuroscience from the University of Paris-Saclay. His PhD is about cellular trafficking and membrane nanodomains in neurons, requiring the use of super-resolution microscopy, such as STED and single molecule localization microscopy.

David Boulet is the technical manager of the primary culture service at the NeurImag facility within the Institut de Psychiatrie et Neurosciences de Paris. He holds a master’s degree in cell biology and integrative neuroscience from Sorbonne University. Within the facility, he provides a unique offer of biological sample preparation ranging from cells to tissues. He is also working on several projects that involved live imaging and high resolution microscopy using STED and SMLM.

Lydia Danglot is a senior researcher at INSERM and scientific director of the NeurImag imaging facility at the Institut de Psychiatrie et Neurosciences de Paris. She received her BS degree in biochemistry from the University Pierre and Marie Curie in 1999 and her PhD in neuroscience in 2004. She is the author of 35 journal papers and 4 reviews. Her current research aims to understand the mechanisms of mammalian neurological synapse morphogenesis using multi-scale imaging and super-resolution techniques, notably SIM, STED, and single molecule localization microscopy.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Victor Breton, Paul Nazac, David Boulet, and Lydia Danglot "Molecular mapping of neuronal architecture using STORM microscopy and new fluorescent probes for SMLM imaging," Neurophotonics 11(1), 014414 (8 March 2024). https://doi.org/10.1117/1.NPh.11.1.014414
Received: 26 October 2023; Accepted: 31 January 2024; Published: 8 March 2024
Advertisement
Advertisement
KEYWORDS
Microscopy

Proteins

Brain mapping

Plasma

Neurons

Confocal microscopy

Molecules

Back to Top