Open Access
22 February 2024 Fiber-based in vivo imaging: unveiling avenues for exploring mechanisms of synaptic plasticity and neuronal adaptations underlying behavior
Anna Karpova, Ahmed A. A. Aly, Endre Levente Marosi, Sanja Mikulovic
Author Affiliations +
Abstract

In recent decades, various subfields within neuroscience, spanning molecular, cellular, and systemic dimensions, have significantly advanced our understanding of the elaborate molecular and cellular mechanisms that underpin learning, memory, and adaptive behaviors. There have been notable advancements in imaging techniques, particularly in reaching superficial brain structures. This progress has led to their widespread adoption in numerous laboratories. However, essential physiological and cognitive processes, including sensory integration, emotional modulation of motivated behavior, motor regulation, learning, and memory consolidation, are intricately encoded within deeper brain structures. Hence, visualization techniques such as calcium imaging using miniscopes have gained popularity for studying brain activity in unrestrained animals. Despite its utility, miniscope technology is associated with substantial brain tissue damage caused by gradient refractive index lens implantation. Furthermore, its imaging capabilities are primarily confined to the neuronal somata level, thus constraining a comprehensive exploration of subcellular processes underlying adaptive behaviors. Consequently, the trajectory of neuroscience’s future hinges on the development of minimally invasive optical fiber-based endo-microscopes optimized for cellular, subcellular, and molecular imaging within the intricate depths of the brain. In pursuit of this goal, select research groups have invested significant efforts in advancing this technology. In this review, we present a perspective on the potential impact of this innovation on various aspects of neuroscience, enabling the functional exploration of in vivo cellular and subcellular processes that underlie synaptic plasticity and the neuronal adaptations that govern behavior.

1.

Evolution of In Vivo Deep Brain Imaging Technologies

The complexity of the brain demands comprehensive methodologies for exploring the interconnected brain regions and understanding how neuronal communication shapes behavior. Highly polarized neurons in the brain feature a numerous en passant presynaptic boutons [Fig. 1(1)], which are crucial for neurotransmitters and neuromodulators release.1 In addition, they possess postsynaptic sites that extend along their dendrites, often associated with compartmentalized dendritic spines or positioned within the dendritic shaft itself.2,3 These intricate structural elements collectively enable an adaptable mode of cell-to-cell communication known as synaptic transmission.4 The consolidation of diverse forms of memories and the neuronal adaptations underlying behavior are orchestrated by the coordinated engagement of pre- and postsynaptic mechanisms within specific brain circuits.511

Fig. 1

Perspectives on in vivo fiber-based imaging: synaptic neurotransmission, neuronal connectivity, structural plasticity, and membrane dynamics in deep brain structures in awake animals. (1) Schematic overview of imaging of synaptic neurotransmission using cpGFP-based reporters designed to detect various neurotransmitters such as glutamate, acetylcholine, GABA, norepinephrine, dopamine, etc. SVs, synaptic vesicles. (2) Structural imaging of dendritic compartmentalization down to the resolution of a single dendritic spines, as well as imaging of neuronal connectivity while animals are actively engaged in learning tasks or social behavior. (3) A schematic drawing of a CA1 pyramidal neuron, depicting distinct dendritic domains capable of receiving unique synaptic inputs. Proximal-to-distal arrows depict dendritic integration. CA1, cornu ammonis 1 subfield of the hippocampus; SO, stratum oriens; SP, stratum pyramidale; SR, stratum radiatum; SLM, stratum lacunosum moleculare; EC, entorhinal cortex. (4) Imaging of membranous organelle trafficking in vivo in correlation with neuronal activity in awake animals. A, autophagosome; ER, endoplasmic reticulum; MVB, multivesicular bodies; RE, recycling endosome; TGN, trans-Golgi network; SA, spine apparatus; LE, late endosome; GS, Golgi satellites; L, lysosome; EE, early endosome; LE, late endosome. (5) The schematic depicts organelle labeling with an acidification probe within neuronal somata of the VTA and LC for in vivo imaging using dual-color ultra-thin fiber-based imaging technology. (Images of animals were obtained from www.biorender.com.)

NPh_11_S1_S11507_f001.png

Two-photon (2P) microscopy has emerged as a promising tool to monitor neuronal activity in behaving animals.12 This technique has undergone continuous refinements, driven by both software-13 and hardware-based advancements, including adaptive optical correction of aberrations,14,15 three-photon approaches,1618 and the design of red-shifted indicators, which reduce the scattering of the emitted light, further improving imaging depth.19,20 Despite the substantial progress achieved in the past two decades, extending imaging depth remains a challenge. This was initially achieved through the implementation of rod-shaped gradient refractive index (GRIN) lenses, typically ranging in diameter from 500 to 1000  μm. These innovations paved the way for the development of miniscopes, enabling imaging in unrestrained animals.2126 While GRIN lens implantation is commonly described as a minimally invasive imaging technique, it is essential to recognize that it can still cause substantial tissue damage. Consequently, this damage can lead to significant necrosis and gliosis that may persist for weeks after surgery.27 Furthermore, the postsurgical effects have the potential to induce artificial dendritic arbor reorganization.28

In the field of neurophotonics, recent breakthroughs with the implementation of multimode fiber-based imaging techniques,29 allowing imaging through ultra-thin fibers, have yielded significant advancements for the in vivo brain imaging.30 Subsequently, these led to the creation of a remarkable 110  μm thin laser-scanning endo-microscope.31,32 This imaging technology is characterized by the minimal tissue damage, sub-1  μm lateral resolution, three-dimensional random-access capabilities, and multiwavelength detection. Furthermore, the fiber small size allows precise positioning within the region of interest, enabling the assessment of neural activity in various dendritic branches and axonal subregions in behaving animals—a distinct advantage not present in GRIN lens technology.

2.

Combinatorial Imaging of Synaptic Neurotransmission, Neuronal Connectivity, and Structural Changes Underlying Plasticity

Understanding the correlation between adaptive behavior and the interplay of presynaptic release, neuronal connectivity, and structural changes in dendritic branches up to the resolution of single dendritic spines and synaptic boutons in the brains of actively engaged animals is of crucial importance. Recent collaborative efforts among genetic engineers and molecular biologists have provided a large repertoire of advanced molecular probes enabling the assessment of neurotransmission in correlation with structural modifications in synaptic compartments [Fig. 1(1)]. Among these probes, there are several generations of single-wavelength reporters designed to sense glutamate, γ-aminobutyric acid (GABA), norepinephrine, dopamine, etc. The first group includes circularly permuted green fluorescent protein (cpGFP)-based iGluSnFR,3335 a membrane-tethered version based on cp superfolder GFP, sf-iGluSnFR, with specific point mutations enabling high spatial and temporal precision in glutamate detection,36,37 an improved version of sf-iGluSnFR featuring enhanced trafficking to the plasma membrane through the addition of Golgi export- and an endoplasmic reticulum (ER) exit motif sequences,38 and a red glutamate sensor based on cp mApple, R-iGluSnFR1.39 The imaging of inhibitory neurotransmission can be effectively accomplished by employing either iGABASnFR or iGABASnFR.mRubi3.40 Several generations of cpGFP-based norepinephrine sensors, with the latest developments referred to as GRAB-NE3.1 and GRAB-NE2h,41,42 are complemented by a range of dopamine sensors extending the dLight1 family,4347 allowing imaging of modulatory inputs. An approach, utilizing fluorescent protein reconstitution at synaptic sites with probes such as mGRASP4850 dual-eGRASP,51 can effectively label synaptic connections within neuronal circuits. Furthermore, the recent advancement in the SynapShot probe, utilizing dimerization-dependent fluorescent proteins, enables the real-time observation of reversible and bidirectional alterations in synaptic contacts during physiological stimulation and enables real-time monitoring of structural changes in synaptic contacts in the brain of behaving mice.52 In addition, Split Protein HEmispheres for REconstitution (SPHERE) technology for both iGluSnFR and iGABASnFR probes has been recently established enabling separate expression in pre- and postsynaptic neurons parts of splitted iGluSnFR and iGABASnFR for the functional detection of neurotransmitters at the contact sites.53 The latter approach provides the opportunity to simultaneously assess functional synaptic neurotransmission and neuronal connectivity for evaluating circuit-related behaviors. In combination with ultra-thin fiber imaging technology, these approaches would enable in vivo imaging of functional neuronal connectivity within the circuit of interest.

While the structural plasticity of dendritic spines, including spine motility, stability, experience-dependent growth, as well as axon branching, synaptic bouton dynamics, and synaptic structure turnover within cortical and dorsal hippocampal areas has been extensively studied,5460 its longitudinal examination in deep brain regions of behaving animals remains relatively unexplored. Synaptic distribution within a neuron’s dendritic arbor is organized into functional clusters and critical for the segregation of synaptic inputs, the phenomenon, known as “clustered synaptic plasticity.”6165 In conjunction with clustered synaptic plasticity, dendritic compartmentalization entails the division of a neuron’s dendritic tree into discrete segments, each with unique functional properties [Figs. 1(2) and 1(3)].66,67 Furthermore, the application of Ca2+ imaging in deep tissue has the potential to significantly enhance our understanding on multiple aspects of Ca2+ signaling and Ca2+ homeostasis. Beyond serving as a valuable tool for providing precise spatiotemporal resolution of axonal spiking activity,68 imaging of Ca2+ transients delves into various facets of Ca2+ homeostasis, including residual Ca2+ concentration, clearance, decay, and summation. Although these aspects have been extensively studied in vitro,6972 the question of the relevance of presynaptic heterogeneity in ongoing behavior remains an open inquiry. Therefore, imaging of Ca2+ transients within subcellular compartments, such as mossy fiber axons in the memory-related brain region, dentate gyrus, in combination with engram technologies such as mGRASP based c-Fos or Arc-dependent labeling of synaptic engrams,48,50,51 enables the comparison of residual Ca2 dynamics between engram and non-engram mossy fiber boutons during memory retrieval.73 This comparative analysis sheds light on the mechanistic underpinnings of function-dependent processing in these synapses, where the coupling between the action potential and the calcium influx (AP-ICa coupling) is indistinguishable.74

Thus, although the application of ultra-thin endo-microscope technology already extends investigations into synaptic neurotransmission, neuronal connectivity, and structural changes underlying plasticity within deep brain regions, further improvement in this technique is adequate. In the future, this technology should enable in vivo examination of the organization and function of axonal boutons larger than 1  μm, including corticothalamic boutons75 and the mnemonic calyx of Held,76 while also allowing imaging of wider areas of dendritic branches and axonal arbors.

3.

Exploring Membrane Trafficking In Vivo in Correlation with Neuronal Activity and Animal Behavior

The ultra-thin laser-scanning endo-microscope allows for subcellular membranous compartment resolution, including vesicular structures of the size of lysosomes.31 This offers new prospects for in vivo subcellular imaging, encompassing multiple membranous compartments, their functional attributes, and trafficking patterns in response to neuronal activity. A subset of membrane trafficking processes includes trafficking of organelles to specific locations within neurons. This ensures the timely delivery of essential building blocks, proteins, and protein complexes while also facilitating a delivery of post-translationally modified transmembrane proteins and receptors to the neuron’s plasma membrane through outward biosynthetic-secretory membrane trafficking.7783 The trafficking of double-membrane autophagosomes84,85 and amphisomes,8689 which subsequently fuse with lysosomes, plays a critical role in maintaining neuronal proteostasis and is integral to the degradation of unwanted cellular cargos such as misfolded proteins and even damaged organelles.90 While there is a continuous exchange of membranes taking place, specific resident proteins or peptides are efficiently employed to target fluorescent proteins to particular organelles and allow for their visualization in living cells [Fig. 1(4)]. Widely used example probes include the GFP-KDEL probe, utilized for ER labeling,91 KFERQ-sequence fused to Dendra for imaging of chaperone-mediated autophagy,92 YFP-ERGIC-53, a marker indicating ER-Golgi Intermediate Compartment (ERGIC),93,94 GM130, which labels somatic Golgi,77 TGN38 for Trans-Golgi-Network labeling,83 the pGolt-mCherry probe for Golgi satellites labeling,81,82,95 Rab5 for early- and Rab7 for late-endosome labeling,96 the autophagy-related protein (ATG) ATG8/LC3 for labeling a subset of autophagosomes,97 and LAMP1 and LAMP2 for lysosome labeling.97101 Furthermore, alterations in organelle function can be monitored in various animal models of neurodevelopmental and neurodegenerative disorders, particularly in the deep brain regions such as ventral tegmental area (VTA) and locus coeruleus (LC), which are particularly vulnerable to neurodegeneration. For example, the tandem mRFP-GFP fusion with LC3, serving as an acidification probe, differentiates non-acidified autophagosomes labeled with the dual signal from acidified autolysosomes marked with mRFP due to their low pH environment.102 The combination of dual color ultra-thin fiber-based imaging technology and the use of the acidification probe may provide insightful information on alteration in neuronal proteostasis in these brain regions underlying neurodegeneration [Fig. 1(5)]. Moreover, the scientific community has access to diverse mouse models expressing specific organelle markers.82,92,103,104 Thus, circuit-specific organelle labeling approach, when combined with ultra-thin fiber-based imaging technology, provides a novel perspective for visualizing membrane trafficking and specific organelle functions in deep brain regions of both healthy and diseased animals.

4.

Limitations and Future Developments

In this review, we have centered our discussion on elucidating the perspectives and emerging opportunities afforded by the development of ultra-thin fiber imaging technology. Although the challenge of achieving longitudinal imaging with the immersible endo-microscope remains unresolved, the potential to enable the continuous observation of neural connections across consecutive days, may be enabled by custom connector solutions or exploitations of implantable thin-wall guiding glass tubes,23 thereby providing valuable insights into the dynamics of learning or behavioral assessments.

The integration of GRIN lenses into miniscope applications has marked a significant advancement in neuroscience over the past decade, enabling the observation of brain activity in real-time during ongoing behavior. Looking ahead, we envision a synergistic relationship between these two technologies, tailored to the specific scientific inquiries. While a GRIN lens boasts a larger size and an expanded field of view, the microendoscope is constrained to 100  μm in diameter, making it better suited for very deep brain structures. Furthermore, the GRIN lens demonstrates compatibility with 2P excitation, facilitating exploration of deeper regions at the GRIN lens output. In contrast, the multimodal fiber endoscope is currently limited to one-photon excitation compatibility. While the microendoscope requires a digital micromirror device and calibration, making it currently more challenging to employ, ongoing developments by the growing scientific community holds promise in developing more user-friendly solutions in the near future. The potential complementarity of GRIN lenses and microendoscope technologies underscores their collective capability to address diverse scientific questions. In summary, the utilization of this imaging technology in freely moving animals introduce a transformative advancement, opening up new possibilities for the investigating cellular, subcellular, and molecular processes.

Disclosures

The authors declare that they have no conflict of interest.

Code and Data Availability

Data sharing is not applicable to this article, as no new data were created or analyzed.

Author Contributions

A.K. and S.M. designed the review. A.K. and S.M. acquired funding. A.K. prepared the figure. A.K., A.A.A.A., E.L.M., and S.M. wrote the paper and all authors commented and revised the manuscript.

Acknowledgments

A.K. is supported by Deutsche Forschungsgemeinschaft (DFG German Research Foundation); Award no(s). FOR5228 RP6 447288260, CRC 1436 TP A02, 42589999, and the “Special project” awarded by Leibniz Institute for Neurobiology (LIN); S.M. is supported by the Leibniz Best Minds Start-Up Grant; Award no. J143/2022, German Center for Psychological Health (DZPG), Award no. Project 01EE2305E, and the Leibniz Institute for Neurobiology. Elements related to in vivo experiments in Fig. 1 were created with BioRender.com.

References

1. 

G. Liu and R. Tsien, “Properties of synaptic transmission at single hippocampal synaptic boutons,” Nature, 375 404 –408 https://doi.org/10.1038/375404a0 (1995). Google Scholar

2. 

K. E. Sorra and K. M. Harris, “Overview on the structure, composition, function, development, and plasticity of hippocampal dendritic spines,” Hippocampus, 10 (5), 501 –511 https://doi.org/10.1002/1098-1063(2000)10:5<501::AID-HIPO1>3.0.CO;2-T (2000). Google Scholar

3. 

H. Hering and M. Sheng, “Dendritic spines: structure, dynamics, and regulation,” Nat. Rev. Neurosci., 2 880 –888 https://doi.org/10.1038/35104061 NRNAAN 1471-003X (2001). Google Scholar

4. 

T. M. Jessell and E. R. Kandel, “Synaptic transmission: a bidirectional and self-modifiable form of cell-cell communication,” Cell, 72 (Suppl), 1 –30 https://doi.org/10.1016/S0092-8674(05)80025-X CELLB5 0092-8674 (1993). Google Scholar

5. 

D. F. Marrone and T. L. Petit, “The role of synaptic morphology in neural plasticity: structural interactions underlying synaptic power,” Brain Res. Rev., 38 (3), 291 –308 https://doi.org/10.1016/S0165-0173(01)00147-3 BRERD2 0165-0173 (2002). Google Scholar

6. 

K. M. Harris, J. C. Fiala and L. Ostroff, “Structural changes at dendritic spine synapses during long- term potentiation,” Philos. Trans. R. Soc. Lond. B Biol. Sci., 358 (1432), 745 –748 https://doi.org/10.1098/rstb.2002.1254 (2003). Google Scholar

7. 

A. Holtmaat et al., “Experience-dependent and cell-type-specific spine growth in the neocortex,” Nature, 441 (7096), 979 –983 https://doi.org/10.1038/nature04783 (2006). Google Scholar

8. 

A. Citri and R. Malenka, “Synaptic plasticity: multiple forms, functions, and mechanisms,” Neuropsychopharmacology, 33 18 –41 https://doi.org/10.1038/sj.npp.1301559 NEROEW 0893-133X (2008). Google Scholar

9. 

J. N. Bourne and K. M. Harris, “Nanoscale analysis of structural synaptic plasticity,” Curr. Opin. Neurobiol., 22 372 –382 https://doi.org/10.1016/j.conb.2011.10.019 COPUEN 0959-4388 (2012). Google Scholar

10. 

N. Vitureira and Y. Goda, “Cell biology in neuroscience: the interplay between hebbian and homeostatic synaptic plasticity,” J. Cell Biol., 203 (2), 175 –186 https://doi.org/10.1083/jcb.201306030 JCLBA3 0021-9525 (2013). Google Scholar

11. 

F. Nadim and D. Bucher, “Neuromodulation of neurons and synapses,” Curr. Opin. Neurobiol., 29 48 –56 https://doi.org/10.1016/j.conb.2014.05.003 COPUEN 0959-4388 (2014). Google Scholar

12. 

K. Svoboda and R. Yasuda, “Principles of two-photon excitation microscopy and its applications to neuroscience,” Neuron, 50 (6), 823 –839 https://doi.org/10.1016/j.neuron.2006.05.019 NERNET 0896-6273 (2006). Google Scholar

13. 

X. Chen et al., “LOTOS-based two-photon calcium imaging of dendritic spines in vivo,” Nat. Prot., 7 (10), 1818 –1829 https://doi.org/10.1038/nprot.2012.106 1754-2189 (2012). Google Scholar

14. 

N. Ji, D. E. Milkie and E. Betzig, “Adaptive optics via pupil segmentation for high-resolution imaging in biological tissues,” Nat. Methods, 7 (2), 141 –147 https://doi.org/10.1038/nmeth.1411 1548-7091 (2010). Google Scholar

15. 

N. Ji, T. R. Sato and E. Betzig, “Characterization and adaptive optical correction of aberrations during in vivo imaging in the mouse cortex,” PNAS, 109 (1), 22 –27 https://doi.org/10.1073/pnas.1109202108 (2012). Google Scholar

16. 

N. G. Horton et al., “In vivo three-photon microscopy of subcortical structures within an intact mouse brain,” Nat. photonics, 7 (3), 205 –209 https://doi.org/10.1038/nphoton.2012.336 NPAHBY 1749-4885 (2013). Google Scholar

17. 

D. G. Ouzounov et al., “In vivo three-photon imaging of activity of GCaMP6-labeled neurons deep in intact mouse brain,” Nat. Methods, 14 (4), 388 –390 https://doi.org/10.1038/nmeth.4183 1548-7091 (2017). Google Scholar

18. 

T. Wang et al., “Quantitative analysis of 1300-nm three-photon calcium imaging in the mouse brain,” eLife, 9 e53205 https://doi.org/10.7554/eLife.53205 (2020). Google Scholar

19. 

C. H. Tischbirek, A. Birkner and A. Konnerth, “In vivo deep two-photon imaging of neural circuits with the fluorescent Ca2+ indicator Cal-590,” J. Physiol., 595 (10), 3097 –3105 https://doi.org/10.1113/JP272790 JPHYA7 0022-3751 (2017). Google Scholar

20. 

W. Wang, C. K. Kim and A. Y. Ting, “Molecular tools for imaging and recording neuronal activity,” Nat. Chem. Biol., 15 (2), 101 –110 https://doi.org/10.1038/s41589-018-0207-0 1552-4450 (2019). Google Scholar

21. 

F. Helmchen et al., “A miniature head-mounted two-photon microscope: high-resolution brain imaging in freely moving animals,” Neuron, 31 (6), 903 –912 https://doi.org/10.1016/S0896-6273(01)00421-4 NERNET 0896-6273 (2001). Google Scholar

22. 

J. C. Jung and M. J. Schnitzer, “Multiphoton endoscopy,” Opt. Lett., 28 (11), 902 –904 https://doi.org/10.1364/OL.28.000902 OPLEDP 0146-9592 (2003). Google Scholar

23. 

R. P. J. Barretto et al., “Time-lapse imaging of disease progression in deep brain areas using fluorescence microendoscopy,” Nat. Med., 17 (2), 223 –228 https://doi.org/10.1038/nm.2292 1078-8956 (2011). Google Scholar

24. 

C. Wang and N. Ji, “Characterization and improvement of three-dimensional imaging performance of GRIN-lens-based two-photon fluorescence endomicroscopes with adaptive optics,” Opt. Express, 21 27142 –27154 https://doi.org/10.1364/OE.21.027142 OPEXFF 1094-4087 (2013). Google Scholar

25. 

S. L. Resendez et al., “Visualization of cortical, subcortical and deep brain neural circuit dynamics during naturalistic mammalian behavior with head-mounted microscopes and chronically implanted lenses,” Nat. Protoc., 11 566 –597 https://doi.org/10.1038/nprot.2016.021 1754-2189 (2016). Google Scholar

26. 

A. Antonini et al., “Extended field-of-view ultrathin microendoscopes for high-resolution two-photon imaging with minimal invasiveness,” eLife, 9 e58882 https://doi.org/10.7554/eLife.58882 (2020). Google Scholar

27. 

M. E. Bocarsly et al., “Minimally invasive microendoscopy system for in vivo functional imaging of deep nuclei in the mouse brain,” Biomed. Opt. Express, 6 (11), 4546 –4556 https://doi.org/10.1364/BOE.6.004546 BOEICL 2156-7085 (2015). Google Scholar

28. 

H. T. Xu, F. Pan and G. Yang, “Choice of cranial window type for in vivo imaging affects dendritic spine turnover in the cortex,” Nat. Neurosci., 10 549 –551 https://doi.org/10.1038/nn1883 NANEFN 1097-6256 (2007). Google Scholar

29. 

E. E. Morales-Delgado, D. Psaltis and C. Moser, “Two-photon imaging through a multimode fiber,” Opt. Express, 23 (25), 32158 –32170 https://doi.org/10.1364/OE.23.032158 OPEXFF 1094-4087 (2015). Google Scholar

30. 

S. A. Vasquez-Lopez et al., “Subcellular spatial resolution achieved for deep-brain imaging in vivo using a minimally invasive multimode fiber,” Light Sci. Appl., 7 110 https://doi.org/10.1038/s41377-018-0111-0 (2018). Google Scholar

31. 

J. Stibůrek et al., “110  μm thin endo-microscope for deep-brain in vivo observations of neuronal connectivity, activity, and blood flow dynamics,” Nat. Commun., 14 (1), 1897 https://doi.org/10.1038/s41467-023-36889-z NCAOBW 2041-1723 (2023). Google Scholar

32. 

H. Cao et al., “Controlling light propagation in multimode fibres for imaging, spectroscopy, and beyond,” Adv. Opt. Photonics, 15 (2), 524 –612 https://doi.org/10.1364/AOP.484298 AOPAC7 1943-8206 (2023). Google Scholar

33. 

J. S. Marvin et al., “An optimized fluorescent probe for visualizing glutamate neurotransmission,” Nat. Methods, 10 (2), 162 –170 https://doi.org/10.1038/nmeth.2333 1548-7091 (2013). Google Scholar

34. 

L. Madisen et al., “Transgenic mice for intersectional targeting of neural sensors and effectors with high specificity and performance,” Neuron, 85 (5), 942 –958 https://doi.org/10.1016/j.neuron.2015.02.022 NERNET 0896-6273 (2015). Google Scholar

35. 

N. Helassa et al., “Ultrafast glutamate sensors resolve high-frequency release at Schaffer collateral synapses,” Proc. Natl. Acad. Sci. U. S. A., 115 (21), 5594 –5599 https://doi.org/10.1073/pnas.1720648115 (2018). Google Scholar

36. 

J. D. Pédelacq et al., “Engineering and characterization of a superfolder green fluorescent protein,” Nat. Biotechnol., 24 (1), 79 –88 https://doi.org/10.1038/nbt1172 NABIF9 1087-0156 (2006). Google Scholar

37. 

J. S. Marvin et al., “Stability, affinity, and chromatic variants of the glutamate sensor iGluSnFR,” Nat. Methods, 15 (11), 936 –939 https://doi.org/10.1038/s41592-018-0171-3 1548-7091 (2018). Google Scholar

38. 

J. D. Vevea and E. R. Chapman, “Acute disruption of the synaptic vesicle membrane protein Synaptotagmin 1 using knockoff in mouse hippocampal neurons,” eLife, 9 e56469 https://doi.org/10.7554/eLife.56469 (2020). Google Scholar

39. 

J. Wu et al., “Genetically encoded glutamate indicators with altered color and topology,” ACS Chem. Biol., 13 (7), 1832 –1837 https://doi.org/10.1021/acschembio.7b01085 (2018). Google Scholar

40. 

J. S. Marvin et al., “A genetically encoded fluorescent sensor for in vivo imaging of GABA,” Nat. Methods, 16 (8), 763 –770 https://doi.org/10.1038/s41592-019-0471-2 1548-7091 (2019). Google Scholar

41. 

J. Feng et al., “A genetically encoded fluorescent sensor for rapid and specific in vivo detection of norepinephrine,” Neuron, 102 (4), 745 –761.e8 https://doi.org/10.1016/j.neuron.2019.02.037 NERNET 0896-6273 (2019). Google Scholar

42. 

J. Feng et al., “Monitoring norepinephrine release in vivo using next-generation GRABNE sensors,” (2023). https://doi.org/10.1101/2023.06.22.546075 Google Scholar

43. 

T. Patriarchi et al., “Ultrafast neuronal imaging of dopamine dynamics with designed genetically encoded sensors,” Science, 360 (6396), eaat4422 https://doi.org/10.1126/science.aat4422 SCIEAS 0036-8075 (2018). Google Scholar

44. 

T. Patriarchi et al., “An expanded palette of dopamine sensors for multiplex imaging in vivo,” Nat. Methods, 17 (11), 1147 –1155 https://doi.org/10.1038/s41592-020-0936-3 1548-7091 (2020). Google Scholar

45. 

F. Sun et al., “A genetically encoded fluorescent sensor enables rapid and specific detection of dopamine in flies, fish, and mice,” Cell, 174 (2), 481 –496.e19 https://doi.org/10.1016/j.cell.2018.06.042 CELLB5 0092-8674 (2018). Google Scholar

46. 

B. Dai et al., “Responses and functions of dopamine in nucleus accumbens core during social behaviors,” Cell Rep., 40 (8), 111246 https://doi.org/10.1016/j.celrep.2022.111246 (2022). Google Scholar

47. 

D. Ino et al., “A fluorescent sensor for real-time measurement of extracellular oxytocin dynamics in the brain,” Nat. Methods, 19 (10), 1286 –1294 https://doi.org/10.1038/s41592-022-01597-x 1548-7091 (2022). Google Scholar

48. 

L. G. Reijmers et al., “Localization of a stable neural correlate of associative memory,” Science, 317 (5842), 1230 –1233 https://doi.org/10.1126/science.1143839 SCIEAS 0036-8075 (2007). Google Scholar

49. 

J. Kim et al., “mGRASP enables mapping mammalian synaptic connectivity with light microscopy,” Nat. Methods, 9 (1), 96 –102 https://doi.org/10.1038/nmeth.1784 1548-7091 (2011). Google Scholar

50. 

B. K. Murthy et al., “Arc-driven mGRASP highlights CA1 to CA3 synaptic engrams,” Front. Behav. Neurosci., 16 1072571 https://doi.org/10.3389/fnbeh.2022.1072571 (2023). Google Scholar

51. 

J. H. Choi et al., “Interregional synaptic maps among engram cells underlie memory formation,” Science, 360 (6387), 430 –435 https://doi.org/10.1126/science.aas9204 SCIEAS 0036-8075 (2018). Google Scholar

52. 

S. Son et al., “Real-time visualization of structural dynamics of synapses in live cells in vivo,” Nat. Methods, https://doi.org/10.1038/s41592-023-02122-4 (2024). Google Scholar

53. 

Y. Shindo et al., “Genetically encoded sensors for analysing neurotransmission among synaptically-connected neurons,” (2022). https://doi.org/10.1101/2022.04.03.486903 Google Scholar

54. 

Y. Zuo et al., “Development of long-term dendritic spine stability in diverse regions of cerebral cortex,” Neuron, 46 (2), 181 –189 https://doi.org/10.1016/j.neuron.2005.04.001 NERNET 0896-6273 (2005). Google Scholar

55. 

D. D. Stettler et al., “Axons and synaptic boutons are highly dynamic in adult visual cortex,” Neuron, 49 (6), 877 –887 https://doi.org/10.1016/j.neuron.2006.02.018 NERNET 0896-6273 (2006). Google Scholar

56. 

V. De Paola et al., “Cell type-specific structural plasticity of axonal branches and boutons in the adult neocortex,” Neuron, 49 (6), 861 –875 https://doi.org/10.1016/j.neuron.2006.02.017 NERNET 0896-6273 (2006). Google Scholar

57. 

A. Holtmaat and K. Svoboda, “Experience-dependent structural synaptic plasticity in the mammalian brain,” Nat. Rev. Neurosci., 10 (9), 647 –658 https://doi.org/10.1038/nrn2699 NRNAAN 1471-003X (2006). Google Scholar

58. 

M. Fu et al., “Repetitive motor learning induces coordinated formation of clustered dendritic spines in vivo,” Nature, 483 (7387), 92 –95 https://doi.org/10.1038/nature10844 (2012). Google Scholar

59. 

A. Attardo, J. E. Fitzgerald and M. J. Schnitzer, “Impermanence of dendritic spines in live adult CA1 hippocampus,” Nature, 523 (7562), 592 –596 https://doi.org/10.1038/nature14467 (2015). Google Scholar

60. 

N. G. Hedrick et al., “Learning binds new inputs into functional synaptic clusters via spinogenesis,” Nat. Neurosci., 25 726 –737 https://doi.org/10.1038/s41593-022-01086-6 NANEFN 1097-6256 (2022). Google Scholar

61. 

N. Govindarajan et al., “Sodium butyrate improves memory function in an Alzheimer’s disease mouse model when administered at an advanced stage of disease progression,” J. Alzheimers Dis., 26 (1), 187 –197 https://doi.org/10.3233/JAD-2011-110080 (2011). Google Scholar

62. 

H. Makino and R. Malinow, “Compartmentalized versus global synaptic plasticity on dendrites controlled by experience,” Neuron, 72 (6), 1001 –1011 https://doi.org/10.1016/j.neuron.2011.09.036 NERNET 0896-6273 (2011). Google Scholar

63. 

T. Kleindienst et al., “Activity-dependent clustering of functional synaptic inputs on developing hippocampal dendrites,” Neuron, 72 (6), 1012 –1024 https://doi.org/10.1016/j.neuron.2011.10.015 NERNET 0896-6273 (2011). Google Scholar

64. 

J. Winnubst and C. Lohmann, “Synaptic clustering during development and learning: the why, when, and how,” Front. Mol. Neurosci., 5 70 https://doi.org/10.3389/fnmol.2012.00070 (2012). Google Scholar

65. 

G. Kastellakis et al., “Synaptic clustering within dendrites: an emerging theory of memory formation,” Prog. Neurobiol., 126 19 –35 https://doi.org/10.1016/j.pneurobio.2014.12.002 PGNBA5 0301-0082 (2015). Google Scholar

66. 

L. Godenzini, A. S. Shai and L. M. Palmer, “Dendritic compartmentalization of learning-related plasticity,” eNeuro, 9 (3), ENEURO.0060-22.2022 https://doi.org/10.1523/ENEURO.0060-22.2022 (2022). Google Scholar

67. 

O. Gemin et al., “Unique properties of dually innervated dendritic spines in pyramidal neurons of the somatosensory cortex uncovered by 3D correlative light and electron microscopy,” PLoS Biol., 19 (8), e3001375 https://doi.org/10.1371/journal.pbio.3001375 (2021). Google Scholar

68. 

C. D. Wilms and M. Häusser, “Reading out a spatiotemporal population code by imaging neighboring parallel fiber axons in vivo,” Nat. Commun., 6 6464 https://doi.org/10.1038/ncomms7464 NCAOBW 2041-1723 (2015). Google Scholar

69. 

W. G. Regehr, K. R. Delaney and D. W. Tank, “The role of presynaptic calcium in short-term enhancement at the hippocampal mossy fiber synapse,” J. Neurosci., 14 (2), 523 –537 https://doi.org/10.1523/JNEUROSCI.14-02-00523.1994 JNRSDS 0270-6474 (1994). Google Scholar

70. 

S. D. Brenowitz and W. G. Regehr, “Reliability and heterogeneity of calcium signaling at single presynaptic boutons of cerebellar granule cells,” J. Neurosci., 27 (30), 7888 –7898 https://doi.org/10.1523/JNEUROSCI.1064-07.2007 JNRSDS 0270-6474 (2007). Google Scholar

71. 

W. Zhang and D. J. Linden, “Calcium influx measured at single presynaptic boutons of cerebellar granule cell ascending axons and parallel fibers,” Cerebellum, 11 (1), 121 –131 https://doi.org/10.1007/s12311-009-0151-3 (2012). Google Scholar

72. 

S. Chamberland et al., “Action potential counting at giant mossy fiber terminals gates information transfer in the hippocampus,” PNAS, 115 (28), 7434 –7439 https://doi.org/10.1073/pnas.1720659115 (2018). Google Scholar

73. 

G. J. Broussard et al., “In vivo measurement of afferent activity with axon-specific calcium imaging,” Nat. Neurosci., 21 1272 –1280 https://doi.org/10.1038/s41593-018-0211-4 NANEFN 1097-6256 (2018). Google Scholar

74. 

E. L. Marosi et al., “Similar presynaptic action potential-calcium influx coupling in two types of large mossy fiber terminals innervating CA3 pyramidal cells,” eNeuro, 10 (2), ENEURO.0017-23.2023 https://doi.org/10.1523/ENEURO.0017-23.2023 (2023). Google Scholar

75. 

N. Hádinger et al., “Region-selective control of the thalamic reticular nucleus via cortical layer 5 pyramidal cells,” Nat. Neurosci., 26 116 –130 https://doi.org/10.1038/s41593-022-01217-z NANEFN 1097-6256 (2023). Google Scholar

76. 

P. H. Smith et al., “Projections of physiologically characterized globular bushy cell axons from the cochlear nucleus of the cat,” J. Comp. Neurol., 304 (3), 387 –407 https://doi.org/10.1002/cne.903040305 JCNEAM 0021-9967 (1991). Google Scholar

77. 

A. C. Horton et al., “Polarized secretory trafficking directs cargo for asymmetric dendrite growth and morphogenesis,” Neuron, 48 (5), 757 –771 https://doi.org/10.1016/j.neuron.2005.11.005 NERNET 0896-6273 (2005). Google Scholar

78. 

T. Wang et al., “Local zones of endoplasmic reticulum complexity confine cargo in neuronal dendrites,” Cell, 148 (1–2), 309 –321 https://doi.org/10.1016/j.cell.2011.11.056 CELLB5 0092-8674 (2012). Google Scholar

79. 

Y. Wu et al., “Contacts between the endoplasmic reticulum and other membranes in neurons,” Proc. Natl. Acad. Sci. U. S. A., 114 (24), E4859 –E4867 https://doi.org/10.1073/pnas.1701078114 (2017). Google Scholar

80. 

A. Perez-Alvarez et al., “Freeze-frame imaging of synaptic activity using SynTagMA,” Nat. Commun., 11 (1), 2464 https://doi.org/10.1038/s41467-020-16315-4 NCAOBW 2041-1723 (2020). Google Scholar

81. 

M. Mikhaylova et al., “A dendritic Golgi satellite between ERGIC and retromer,” Cell Rep., 14 (2), 189 –199 https://doi.org/10.1016/j.celrep.2015.12.024 (2016). Google Scholar

82. 

M. Andres-Alonso et al., “Golgi satellites are essential forpolysialylation of NCAM and expression of LTP at distal synapses,” Cell Rep., 42 (7), 112692 https://doi.org/10.1016/j.celrep.2023.112692 (2023). Google Scholar

83. 

P. P. Y. Lie et al., “Post-Golgi carriers, not lysosomes, confer lysosomal properties to pre-degradative organelles in normal and dystrophic axons,” Cell Rep., 35 (4), 109034 https://doi.org/10.1016/j.celrep.2021.109034 (2021). Google Scholar

84. 

S. E. Hill and D. A. Colón-Ramos, “The journey of the synaptic autophagosome: a cell biological perspective,” Neuron, 105 (6), 961 –973 https://doi.org/10.1016/j.neuron.2020.01.018 NERNET 0896-6273 (2020). Google Scholar

85. 

J. Ye and M. Zheng, “Autophagosome trafficking,” Adv. Exp. Med. Biol., 1208 67 –77 https://doi.org/10.1007/978-981-16-2830-6_5 AEMBAP 0065-2598 (2021). Google Scholar

86. 

X. T. Cheng et al., “Axonal autophagosomes recruit dynein for retrograde transport through fusion with late endosomes,” J. Cell Biol., 209 (3), 377 –386 https://doi.org/10.1083/jcb.201412046 JCLBA3 0021-9525 (2015). Google Scholar

87. 

M. Andres-Alonso et al., “SIPA1L2 controls trafficking and local signaling of TrkB-containing amphisomes at presynaptic terminals,” Nat. Commun., 10 (1), 5448 https://doi.org/10.1038/s41467-019-13224-z NCAOBW 2041-1723 (2019). Google Scholar

88. 

M. Andres-Alonso, M. R. Kreutz and A. Karpova, “Autophagy and the endolysosomal system in presynaptic function,” Cell Mol. Life Sci., 78 (6), 2621 –2639 https://doi.org/10.1007/s00018-020-03722-5 CMLSFI 1420-9071 (2021). Google Scholar

89. 

K. M. Grochowska et al., “The needs of a synapse—how local organelles serve synaptic proteostasis,” EMBO J., 41 e110057 https://doi.org/10.15252/embj.2021110057 EMJODG 0261-4189 (2022). Google Scholar

90. 

N. Mizushima, T. Yoshimori and Y. Ohsumi, “The role of ATG proteins in autophagosome formation,” Annu. Rev. Cell Dev. Biol., 27 107 –132 https://doi.org/10.1146/annurev-cellbio-092910-154005 ARDBF8 1081-0706 (2011). Google Scholar

91. 

A. Perez-Alvarez et al., “Endoplasmic reticulum visits highly active spines and prevents runaway potentiation of synapses,” Nat. Commun., 11 5083 https://doi.org/10.1038/s41467-020-18889-5 NCAOBW 2041-1723 (2020). Google Scholar

92. 

S. Dong et al., “Monitoring spatiotemporal changes in chaperone-mediated autophagy in vivo,” Nat. Commun., 11 645 https://doi.org/10.1038/s41467-019-14164-4 NCAOBW 2041-1723 (2020). Google Scholar

93. 

C. M. Gürth et al., “Synaptic activity and strength are reflected by changes in the post-synaptic secretory pathway,” Sci. Rep., 10 (1), 20576 https://doi.org/10.1038/s41598-020-77260-2 SRCEC3 2045-2322 (2020). Google Scholar

94. 

M. de Los Ángeles Juricic Urzúa et al., “The dendritic ERGIC: microtubule and actin cytoskeletons participate in stop- and -go movement of mobile carriers between stable structures,” Traffic, 23 (3), 174 –187 https://doi.org/10.1111/tra.12832 TRAQA4 1398-9219 (2022). Google Scholar

95. 

A. P. Govind et al., “Activity-dependent Golgi satellite formation in dendrites reshapes the neuronal surface glycoproteome,” eLife, 10 e68910 https://doi.org/10.7554/eLife.68910 (2021). Google Scholar

96. 

K. Deinhardt et al., “Rab5 and Rab7 control endocytic sorting along the axonal retrograde transport pathway,” Neuron, 52 (2), 293 –305 https://doi.org/10.1016/j.neuron.2006.08.018 NERNET 0896-6273 (2006). Google Scholar

97. 

S. Maday and E. L. F. Holzbaur, “Autophagosome biogenesis in primary neurons follows an ordered and spatially regulated pathway,” Dev. Cell, 30 (1), 71 –85 https://doi.org/10.1016/j.devcel.2014.06.001 1534-5807 (2014). Google Scholar

98. 

C. A. Boecker et al., “ToolBox: live imaging of intracellular organelle transport in induced pluripotent stem cell-derived neurons,” Traffic, 21 (1), 138 –155 https://doi.org/10.1111/tra.12701 TRAQA4 1398-9219 (2020). Google Scholar

99. 

X. T. Cheng et al., “Characterization of LAMP1-labeled nondegradative lysosomal and endocytic compartments in neurons,” J. Cell Biol., 217 (9), 3127 –3139 https://doi.org/10.1083/jcb.201711083 JCLBA3 0021-9525 (2018). Google Scholar

100. 

S. E. Cason et al., “Sequential dynein effectors regulate axonal autophagosome motility in a maturation-dependent pathway,” J. Cell Biol., 220 (7), e202010179 https://doi.org/10.1083/jcb.202010179 JCLBA3 0021-9525 (2021). Google Scholar

101. 

K. M. Grochowska et al., “Chaperone-mediated autophagy in neuronal dendrites utilizes activity- dependent lysosomal exocytosis for protein disposal,” Cell Rep., 42 (8), 112998 https://doi.org/10.1016/j.celrep.2023.112998 (2023). Google Scholar

102. 

S. Kimura, T. Noda and T. Yoshimori, “Dissection of the autophagosome maturation process by a novel reporter protein, tandem fluorescent-tagged LC3,” Autophagy, 3 (5), 452 –460 https://doi.org/10.4161/auto.4451 (2007). Google Scholar

103. 

N. Mizushima and A. Kuma, “Autophagosomes in GFP-LC3 transgenic mice,” Methods Mol. Biol., 445 119 –124 https://doi.org/10.1007/978-1-59745-157-4_7 (2008). Google Scholar

104. 

N. Mizushima, “Methods for monitoring autophagy using GFP-LC3 transgenic mice,” Methods Enzymol., 452 13 –23 https://doi.org/10.3390/cells6020014 MENZAU 0076-6879 (2009). Google Scholar

Biography

Anna Karpova pursued her diploma studies at Bashkir State University (BSU) in Ufa, specializing in biology and chemistry (excellent). Subsequently, she obtained her PhD in neuroscience through a collaborative project between BSU and Lomonosov State University in Moscow, focusing on the amygdala's circuitry in rodents. In 2003, she moved to Magdeburg and is currently a project leader in the Neuroplasticity Research Group at the Leibniz Institute for Neurobiology. Her research interests include synapse biology, membrane trafficking, and the cellular and molecular mechanisms governing activity-dependent and homeostatic synaptic plasticity, all while utilizing cutting-edge imaging technologies.

Ahmed A. A. Aly obtained his bachelor's degree in biophysics from Ain Shams University in Cairo in 2017. Subsequently, he obtained his master’s degree in neuroscience from Santiago Canals’ Lab at the Instituto de Neurociencias - UMH in Alicante, Spain, in 2019. Currently, his PhD project is supported by funding from the LIN Special Project. Under Anna Karpova’s guidance, his research centers on the in vivo regulation of autophagy processes, employing two-photon microscopy and fiber-based photoconversion methodologies.

Endre Levente Marosi completed the biology program at Eötvös Loránd University, Budapest, specializing in studying the action potential–calcium current coupling in hippocampal mossy fiber terminals through direct axonal patch clamp recordings at the Institute of Experimental Medicine, Budapest. Since October 2020, he has been pursuing a PhD at the Leibniz Institute for Neurobiology, Magdeburg, with a current research emphasis on the medial septal–hippocampal circuitry particularly during locomotion.

Sanja Mikulovic studied biomedical engineering at the Technical University of Vienna and received her doctorate in neuroscience from the University of Uppsala in Sweden. Since her postdoc time at Uppsala University, Karolinska Institute in Stockholm, and DZNE Bonn, she has been focusing on investigating neural networks underlying cognitive and emotional behavior, using combination of electrophysiology and optical techniques. Since January 2021, she has been leading her own group at the Leibniz Institute for Neurobiology in Magdeburg.

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.
Anna Karpova, Ahmed A. A. Aly, Endre Levente Marosi, and Sanja Mikulovic "Fiber-based in vivo imaging: unveiling avenues for exploring mechanisms of synaptic plasticity and neuronal adaptations underlying behavior," Neurophotonics 11(S1), S11507 (22 February 2024). https://doi.org/10.1117/1.NPh.11.S1.S11507
Received: 31 October 2023; Accepted: 23 January 2024; Published: 22 February 2024
Advertisement
Advertisement
KEYWORDS
Brain

In vivo imaging

Neuroimaging

Animals

Calcium

Neuroscience

Biological imaging

Back to Top