Home Synapses: Multitasking Global Players in the Brain
Article Publicly Available

Synapses: Multitasking Global Players in the Brain

  • Joachim H. R. Lübke

    Born 11.04.1956 in Recklinghausen, Germany; 1974–1977: Education as a biological technician; 1981–1982: ‘Begabten-Abitur’ at the Abend-gymnasium Göttingen; 1982–1987: Study of Biology, Georg-August Universität Göttingen; 1987–1991: Diploma and doctoral thesis, Max-Planck-Institut für Biophysikalische Chemie, Abt. Neurobiologie (Prof. Dr. O.-D. Creutzfeldt); 1991–1993: Postdoctoral Fellow (Scolarship of the Royal Society of Science), Dept. of Human Anatomy (Prof. Dr. Ray Guillery, University of Oxford; 1993–1995: ‘Von Helmholtz Stipendiat’ of the BMBF, Anatomisches Institut (Prof. Dr. Michael Frotscher), Albert-Ludwigs Universität Freiburg; 1996: Wolfgang-Bargmann Preis, Anatomische Gesellschaft; 1996: Fachanatom, Anatomische Gesellschaft; 1999: Habilitation in Anatomy and Neuroanatomy, Albert-Ludwigs Universität Freiburg; 2000: Appointment as a C2 Lecturer, Albert-Ludwigs Universität Freiburg; 2003: Tenured-track appointment and Group leader position ‘Cellular Neurobiology’, Institute of Medicine (Prof. Dr. Karl Zilles, Research Centre Jülich GmbH; 2005: APL-Professor in Anatomy and Neuroanatomy, Heinrich-Heine Universität Düsseldorf; Since 2008: Professor W2, Medical Faculty, University Hospital/RWTH Aachen. Current Position: W2-Proffesor and group leader ‘Structure of Synapses’, Institute of Neuroscience and Medicine INM-10, Research Centre Jülich GmbH.

    ORCID logo EMAIL logo
    and Astrid Rollenhagen

    Born 06.11.1959 in Bielefeld, Germany; 1976–1978: Education as a doctor’s assistant; 1980–1982 ‘Abitur’ at the Westfalen Kolleg Bielefeld; 1982–1987: Study of Biology, Universität Bielefeld; 1987–1996: Diploma and doctoral thesis, Universität Bielefeld, Lehrstuhl Verhaltensforschung, AG Prof. Dr. H.-J. Bischof; 1996–1997: Scientific Assistant, Universität Bielefeld, Lehrstuhl Verhaltensforschung, AG Prof. Dr. H.-J. Bischof; 1998–2002: Universität Hamburg, Center for Molecular Neurobiology Hamburg, Abt. Prof. Dr. M. Schachner; 2002–2004: Albert-Ludwigs-Universität Freiburg, Dept. of Anatomy (Prof. Dr. M. Frotscher, AG PD Dr. J.H.R. Lübke); Since 2004: Senior Post-Doc, Group: ‘Structure of Synapses’ (Prof. Dr. J.H.R. Lübke), INM-10, Research Centre Jülich GmbH.

    EMAIL logo
Published/Copyright: December 5, 2019
Become an author with De Gruyter Brill

Abstract

Synapses are key elements in the communication between neurons in any given network of the normal adult, developmental and pathologically altered brain. Synapses are composed of nearly the same structural subelements: a presynaptic terminal containing mitochondria with an ultrastructurally visible density at the pre- and postsynaptic apposition zone. The presynaptic density is composed of a cocktail of various synaptic proteins involved in the binding, priming and docking of synaptic vesicles inducing synaptic transmission. Individual presynaptic terminals (synaptic boutons) contain a couple of hundred up to thousands of synaptic vesicles. The pre- and postsynaptic densities are separated by a synaptic cleft. The postsynaptic density, also containing various synaptic proteins and more importantly various neurotransmitter receptors and their subunits specifically composed and arranged at individual synaptic complexes, reside at the target structures of the presynaptic boutons that could be somata, dendrites, spines or initial segments of axons.

Beside the importance of the network in which synapses are integrated, their individual structural composition critically determines the dynamic properties within a given connection or the computations of the entire network, in particular, the number, size and shape of the active zone, the structural equivalent to a functional neurotransmitter release site, together with the size and organization of the three functionally defined pools of synaptic vesicles, namely the readily releasable, the recycling and the resting pool, are important structural subelements governing the ‘behavior’ of synaptic complexes within a given network such as the cortical column.

In the late last century, neuroscientists started to generate quantitative 3D-models of synaptic boutons and their target structures that is one possible way to correlate structure with function, thus allowing reliable predictions about their function. The re-introduction of electron microscopy (EM) as an important tool achieved by modern high-end, high-resolution transmission-EM, focused ion beam scanning-EM, CRYO-EM and EM-tomography have enormously improved our knowledge about the synaptic organization of the brain not only in various animal species, but also allowed new insights in the ‘microcosms’ of the human brain in health and disease.

Zusammenfassung

Synapsen sind Schlüsselelemente der Kommunikation zwischen Neuronen in jedem beliebigen Netzwerk des normal adulten, sich entwickelnden, bzw. krankhaft veränderten Gehirns. Synapsen sind nahezu aus den gleichen strukturellen Subelementen aufgebaut: einem präsynaptischen Element, welches Mitochondrien und eine ultrastrukturell sichtbare Proteinverdichtung der Membran mit einem Cocktail verschiedener synaptischer Proteine enthält, welche für die Bindung‚ das „Priming“ und das Andocken synaptischer Vesikel verantwortlich sind. Das präsynaptische Terminal (synaptischer Bouton) kann einige hundert bis zu einigen tausend synaptische Vesikel enthalten. Die präsynaptische Seite ist durch den synaptischen Spalt von der postsynaptischen Dichte der Zielstruktur getrennt, die entweder Somata, Dendriten, dendritische „Spines“ oder Axoninitialsegmente darstellen. Die postsynaptische Dichte enthält wiederum spezifische synaptische Proteine, aber noch wichtiger verschiedene Neurotransmitter-Rezeptoren und deren Untereinheiten, die je nach Synapsentyp individuell komponiert und arrangiert sind.

Zum Ende des letzten Jahrhunderts wurde begonnen quantitative 3D-Modelle synaptischer Boutons und deren Zielstrukturen zu generieren, welches eine Möglichkeit darstellt korrelierte Struktur/Funktions-Beziehungen herzustellen. In anderen Worten: erlaubt die strukturelle Komposition von Synapsen verlässliche Voraussagen zu ihrer Funktion?

Die „Wiederentdeckung“ der Elektronmikroskopie (EM) als ein wichtiges Instrument hat mittels hochmoderner, hochauflösender Transmission-EM, der Einführung der „Focused Ion Beam Scanning-EM“ Technologie, die Etablierung von CRYO-EM sowie EM-Tomographie zu einem enormen Erkenntnisgewinn der synaptischen Organisation in verschiedenen Tiermodellen, aber auch zu neuen Erkenntnissen im „Mikrokosmos“ des gesunden und erkrankten menschlichen Gehirns geführt.

Historical background

For centuries, the belief that the structure of the brain and its elements, namely neurons, their dendrites, axons and synapses, and non-neuronal astro- and oligodendrocytes, reflect its function had been a ‘driving force’ for investigations and the source of major discoveries in neuroscience. One of the most important ones was, beside the definition of the neuron, the introduction of the term ‘Synapse’ originally termed more than 100 years ago by Charles Sherrington, a well-known electrophysiologist at that time. Ramon y Cajal, one of the most influential neuroanatomists and the founder of the neuronal doctrine, later adopted this term. The word ‘Synapse’ comes from the Greek synapsis (συνάψις), meaning ‘conjunction’, and from συνάπτεὶν (συν ‘together’) and ἅπτειν (‘to fasten’). This early fundamental discovery and description is even more intriguing because both Sherrington and Ramon y Cajal suggested a direct connection between neurons via these structures, but without ever seeing them.

The introduction of EM and its further development leading nowadays to high-end, fine-scale transmission electron microscopy (TEM), focused ion beam scanning electron microscopy (FIB-SEM), CRYO-EM and EM tomography combined, for example with high-pressure freezing (see Studer et al. 2014; Imig et al. 2014) extended the structural investigations from the light microscopic visible neuron to the subcellular and even the molecular level of the synapse, the elementary building block of any neural networks in the brain. It has to be mentioned thought, that CRYO-EM is more typically suited for the analysis of the structure of proteins at high resolution, rather than subcellular structures.

In addition, modern light microscopic techniques, like Stimulated Emission Depletion (STED) and direct stochastic optical reconstruction microscopy (dSTORM) aimed to visualize synaptic structures at the nanoscopic level, however focused more on the abundance of various synaptic proteins at active zones. Furthermore, the introduction of CRYO-correlative light- and electron microscopy (CRYO-CLEM) allows both fluorescence microscopy as well as three dimensional (3D) CRYO-EM tomography to reveal the ultrastructure of significant target molecules with specific cellular functions at high temporal and spatial resolution (Plitzko et al. 2009). Since then, a wealth of information was obtained that deepened our understanding of synaptic structures, in the developmental and adult brain in health and disease based on studies undertaken in various animal species, including rodents, higher mammals, non-human primates and even humans.

Although synapses had been looked at from different viewpoints, summarized in meanwhile thousands of original publications, reviews and numerous textbooks, a detailed, comprehensive and quantitative knowledge about their morphology is still limited to a relative small number of CNS synapses in different brain regions (Calyx of Held: Rowland et al. 2000; Sätzler et al. 2002; Wimmer et al. 2006; Cochlear bushy cell synapses: Nicol and Walmsley 2002; Climbing fiber synapses: Xu-Friedman et al. 2001; Cerebellar mossy fiber: Xu-Friedman and Regehr 2003; Hippocampal Mossy Fiber Bouton: Chicurel and Harris 1992; Rollenhagen et al. 2007; Synapses in the dentate gyrus: Marrone et al. 2005; Area CA1 synapses: Sorra and Harris 1993; Harris and Sultan 1995; Spacek and Harris 1998; Schikorski and Stevens 1997, 2001; Ribbon synapses in the retina and cochlear: Sikora et al. 2005; Moser et al. 2006; Michanski et al. 2019; Olfactory cortical synapses: Schikorski and Stevens 1999). Such detailed descriptions, however, are required to understand and link structural and functional components of the signal cascades underlying synaptic transmission and plasticity.

An important first step towards an improved understanding of synaptic function were simultaneous patch-clamp recordings from a glutamatergic giant synapse, the so-called Calyx of Held terminating on the principal neurons in the medial nucleus of the trapezoid body in the auditory brainstem by Sakmann, Neher and co-workers (for example see Borst and Sakmann 1996, 1998, 1999; Takahashi et al. 1996; Schneggenburger et al. 1999, Schneggenburger and Neher 2000; for review see also Schneggenburger et al. 2002). However, it turned out that the Calyx of Held is rather the exception than the rule with respect to its synaptic properties perfectly adapted to audition. Hence the investigation of the Calyx of Held synapse strongly suggested that synapses are ‘unique’ entities, in both structural and functional terms.

The second, more central synapse, where paired recording became possible was the mossy fiber bouton-CA3 pyramidal cell synapse in the hippocampus, a synapse involved in learning and memory processes (Geiger and Jonas 2000; Bischofberger and Jonas 2002; Hallermann et al. 2003; Engel and Jonas 2005; Alle and Geiger 2006). The work on this synapse strongly supported and extended the view that synapses are unique in their structural and functional properties. Thus, the dream to create a general ‘model synapse’ for the brain was over.

Nevertheless, the simultaneous recordings from two different CNS synapses and their target structures made it possible for the first time to measure transmitter release under defined internal and external ionic and membrane potential conditions. In addition, the size and time course of action potential evoked Ca2+ influx (Borst and Sakmann 1996, 1998; Bischofberger and Jonas 2002), the occupancy of the putative Ca2+ sensor driving vesicle fusion (Bollmann et al. 2000; Schneggenburger and Neher 2000), the equilibration of intracellular Ca2+ with the endogenous Ca2+ buffer, and the eventual Ca2+-clearance (Helmchen et al. 1997) can be accurately measured. Furthermore, the latency, size and time course of evoked quantal and multiquantal EPSCs (Borst and Sakmann 1996; Silver et al. 2003; Molnar et al. 2016; Holderith et al. 2016; Seeman et al. 2018; Rollenhagen et al. 2018; Vaden et al. 2019; reviewed by Neher 2015; Chamberland and Toth 2016) can be determined. However, there are still steps in the signal cascades that at present can only be simulated (Yamada and Zucker 1992; Bertram et al. 1999; Meinrenken et al. 2002, 2003; Freche et al. 2011). This includes the site, time- and space-dependent build-up and collapse of Ca2+-domains around the pore of Ca2+ channels at a synaptic contact and the buffered diffusion and the subsequent interaction of free Ca2+ with the Ca2+ sensor.

Thus, realistic values of the geometry of synaptic boutons, including the number, size and shape of active zones, and the three functionally defined pools of synaptic vesicles (Rizzoli and Betz 2005), namely the readily releasable (RRP), the recycling (RP) and resting pool are essential for constraining realistic geometrical models of synaptic structures.

On the postsynaptic side, the time course and amplitude of spontaneous and evoked excitatory postsynaptic potentials (EPSCs) were used to infer the characteristics of quantal release (Henze et al. 1997; Silver et al. 2003; Biro et al. 2005; Szabadics et al. 2006; Saviane and Silver 2006; Rollenhagen et al. 2018; Vaden et al. 2019). This interference requires simulations of the transient increase of the glutamate concentration in the synaptic cleft, reversible binding of glutamate to appropriate glutamate receptors and eventual uptake and diffusion of glutamate out of the cleft (Freche et al. 2011). To a large extent, these processes are governed by the geometry of the synaptic cleft and the shape and size of pre- and postsynaptic densities. These parameters can only be estimated from 3D-reconstructions of synaptic structures.

This review will focus on recent findings on the detailed quantitative structural description of the most common type of synaptic bouton in the CNS: excitatory synaptic boutons in different layers of the rodent, non-human primate and human neocortex. Their small size and the great diversity of neurons and synaptic boutons in different layers of a cortical column made this investigation a challenging and difficult task. Such detailed morphological descriptions are useful to directly correlate structure with function of synapses and may therefore explain their different and specific functional performance and computational properties within the network in which they are integrated. On the other hand, such quantitative data provide the basis for numerical and/or MonteCarlo simulations of various synaptic parameters that are still only partially accessible for experiment, at least in the human brain.

Methodological considerations

One possible way to describe synaptic boutons and their target structures in such great detail are either 3D-volume reconstruction based on serial ultrathin sections using TEM (Fig. 1A) or FIB-SEM (Fig. 1B). With the second approach, serial digital EM images were obtained by constant milling a defined area of the sample containing the area of interest by a gallium ion laser beam and subsequent imaging of the block surface (block-face imaging).

From the resulting z-stacks of EM images quantitative 3D-models of synaptic boutons and their prospective target structures can then be generated using different commercially or self-made reconstruction software tools running on high-performance computer systems. Both, TEM and FIB-SEM have advantages, but also disadvantages. Serial sectioning and subsequent TEM examination of ultrathin sections within a series is a very labor-intensive and thus time-consuming process with a comparable low throughput of tissue samples. Secondly, in ultrathin sections, the tilting of the electron beam restricts the area of interest, and during the cutting and imaging process, malformations or distortions or the complete loss of the tissue sample can be a limiting factor. However, the major advantage of using serial ultrathin sections and TEM imaging is their very high quality at high resolution that is required for the detailed analysis of important structural subelements such as the number, size and shape of active zones and the organization and size of the three functionally defined pools of synaptic vesicles (Figs. 1A, 2C, D, 3A-E).

Fig. 1: Comparison of the ultrastructure between TEM and FIB-SEM
A, Low power electron micrograph of the neuropil in the lower part of layer 1 in the human temporal lobe neocortex as visualized with TEM. Note that even at this relatively low EM magnification several synaptic complexes between synaptic boutons and either dendritic shafts or spines, are clearly identifiable.

			B, Low power electron micrograph of the neuropil in layer 4 of the human temporal lobe neocortex taken with FIB-SEM. Scale bar in A and B 1 µm.
			In both electron micrographs, synaptic boutons are given in transparent yellow and postsynaptic structures in transparent blue. Several thick astrocytic (ast) processes are clearly visible.
			Note the differences in the appearance of active zones and synaptic vesicles due to the use of a different EM protocol required for SEM-FIB.
Fig. 1:

Comparison of the ultrastructure between TEM and FIB-SEM

A, Low power electron micrograph of the neuropil in the lower part of layer 1 in the human temporal lobe neocortex as visualized with TEM. Note that even at this relatively low EM magnification several synaptic complexes between synaptic boutons and either dendritic shafts or spines, are clearly identifiable.

B, Low power electron micrograph of the neuropil in layer 4 of the human temporal lobe neocortex taken with FIB-SEM. Scale bar in A and B 1 µm.

In both electron micrographs, synaptic boutons are given in transparent yellow and postsynaptic structures in transparent blue. Several thick astrocytic (ast) processes are clearly visible.

Note the differences in the appearance of active zones and synaptic vesicles due to the use of a different EM protocol required for SEM-FIB.

In contrast, FIB-SEM (Fig. 1B), a relatively new, modern EM technology, allows a much higher throughput of tissue samples because the time and labor-intensive step of serial ultrathin sectioning is no longer required. Secondly, a larger area of interest ~50 by 50 µm can be obtained compared to TEM where the area of interest is limited to ~10–20 by 10–20 µm. Finally, since the surface of the block is milled and polished rather no malformations or distortions are expected thus no or minor alignment processing of adjacent images is required. The major disadvantage, however, are limitations in the resolution of active zones and synaptic vesicles which appear structurally different due to the use of a different EM embedding protocol required for FIB-SEM (Fig. 1B; Movie 1).

In the future, the combination of both TEM and FIB-SEM will be the method of choice to address specific questions and further unravel the ‘microcosms’ of the brain, for example in describing the ‘connectomics’ and synaptic organization of various layers, nuclei and brain regions.

Synaptic boutons in the neocortex of rodents and non-human primates

Meanwhile numerous publications described structural and functional aspects of synaptic transmission and plasticity in different layers mainly in the rodent neocortex using paired or multiple recordings and subsequent morphological analysis of synaptically coupled pairs filled with biocytin or fluorescent dyes during recording (reviewed by Lübke and Feldmeyer 2007; Feldmeyer 2012; Feldmeyer et al. 2013; Qi et al. 2015; Radnikow and Feldmeyer 2018). It has been demonstrated that, beside similarities huge differences exist between intralaminar (synaptic connections within a given layer) and translaminar (synaptic connections across layers) excitatory-excitatory, excitatory-inhibitory and inhibitory-inhibitory synaptic connections with respect to synaptic efficacy, strength, release probability, short-term plasticity and contribution of various neurotransmitter receptors and their subunits, for example different glutamate and GABA receptors. However, these studies aimed to correlate structural with functional properties of a given synaptic connection rather than on the structural composition of individual synaptic contacts.

In contrast, only a few coherent and quantitative structural studies exist for synaptic boutons in the rodent neocortex (Rollenhagen et al. 2015, 2018; Dufour et al. 2016; Bopp et al. 2017; Hsu et al. 2017; Rodriguez-Moreno et al. 2018) and non-human primate neocortex (Anderson and Martin 2006, 2009; Freese and Amaral 2006; Hsu et al. 2017). These studies have demonstrated, for example, layer, region and gender specific differences in the density of synaptic boutons (Alonso-Nanclares et al. 2008). Most strikingly, synaptic boutons beside layer and area-specific differences (see also Rollenhagen 2015, 2018; Bopp et al. 2017; Hsu et al. 2017), differ substantially not only in their shape and size, but even more importantly in the number, size and shape of active zones and in the organization and size of the three pools of synaptic vesicles summarized in Table 1. Interestingly, some structural parameters such as bouton size, pre- and postsynaptic density surface area, content of mitochondria, and synaptic vesicles pools are in some cortical synapses well correlated but in others, no or only a weak correlation between several structural subelements are found (Rollenhagen 2015, 2018; Dufour et al. 2016; Hsu et al. 2017; Bopp et al. 2017; Rodriguez-Moreno et al. 2018). The most striking difference at cortical synaptic boutons is the total pool, and the three functionally defined pools of synaptic vesicles, namely the RRP, the RP and resting pools. Is has to be noted that a structural correlate for the functionally defined pools is not identifiable at the EM level due to the ‘randomly’ distribution of synaptic vesicles within the synaptic bouton. However, an attempt was made to sort synaptic vesicles with respect to their distance from the presynaptic density by a perimeter analysis. This approach allows the identification of synaptic vesicles belonging to one of the functionally defined pools (for criteria see Rizzoli and Betz 2005) and match nearly perfectly with functional estimations of the RRP (Hallermann et al. 2003) and RP (Rollenhagen et al. 2018). However, it is for example, still controversially discussed whether only so-called ‘docked’ vesicles identified at the ultrastructural level (Figs. 3F, 4C inset) already fused at the presynaptic density represent the RRP or also vesicles very close (10–20 nm from the active zone) belong to the RRP (Rollenhagen et al. 2015, 2018; Yakoubi et al. 2019a, b). The same holds true for the RP, how is it defined, and how many vesicles it contains at which distance from the presynaptic density is largely unknown for most of the CNS synapses (but see Rollenhagen et al. 2018). For the role and importance of the resting pool of synaptic vesicles that is thought not to be recruited under ‘normal’ physiological conditions, rather no information is available. However, it has been shown that vesicles from the resting pools can be transferred to the RP and RRP under the control of mitochondria (Verstreken et al. 2005). As a consequence, further experiments using a combination of labeling synaptic vesicles, for example with SM1–43, and high-resolution STED or two-photon laser microcopy using in vitro acute brain slices or neuronal cell cultures can address such questions (Verstreken et al. 2008) in more detail.

Table 1:

Comparison of structural parameters of synaptic boutons in different regions or layers of the rodent, non-human primate and human neocortex

Species

(Strain)

Mouse

(C57/BL6)*

Mouse

(C57/BL6)**

Mouse

(C57/BL6)***

Rat

(Wistar)

Monkey

(Macaca mulatta)***

Human

Region

L4_M1

L4_S1

L4_S1

L2–3_V1

L2–3_FC

L4_S1

L5_S1

L2–3_V1

L2–3_LPFC

L4_TL

L5_TL

Sex

♀ / ♂

♀ / ♂

♀ / ♂

♀ / ♂

♀ / ♂

♀ / ♂

♀ / ♂

Age

60–65 days

2–14 month

90–120 days

5–20 years

20–63 years

Synaptic boutons

Density

16.3 ± 2.7

31.7 ± 4.9

0.93 ± 0.08

1.05 ± 0.13

0.52 ± 0.06

0.48 ± 0.04

2.37x106

Surface area (µm2)

4.67 ± 2.20

3.03 ± 0.71

8.19 ± 2.84

2.50 ± 1.78

6.09 ± 0.85

Volume (µm3)

0.169+ / 0.067-

0.306+ / 0.070-

0.46 ± 0.27

0.10 ± 0.01

0.08 ± 0.03

0.20 ± 0.07

0.38 ± 0.23

0.20 ± 0.04

0.30 ± 0.01

0.16 ± 0.16

0.63 ± 0.17

Mitochondria

Volume (µm3)

0.034+ / 0.00-

0.061+ / 0.008-

0.09 ± 0.06

0.05 ± 0.02

0.07 ± 0.05

0.03 ± 0.04

0.12 ± 0.10

% of the total volume

20.12+ / 0.000-

19.93+ / 11.43-

23.03

20.29

15.09

13.11

12.04

Active zones

Number per bouton

1.3+ / 1.1-

2.1+ / 1.2-

1.6

1.06 ± 0.06

1.12 ± 0.09

1–3

1–2

PreAZ surface area (µm2)

0.18 ± 0.06

0.29 ± 0.19

0.13 ± 0.07

0.23 ± 0.05

PSD surface area (µm2)

0.064+ / 0.56-

0.042+ / 0.039-

0.21 ± 0.11

0.08 ± 0.01

0.07 ± 0.02

0.18 ± 0.06

0.31 ± 0.21

0.08 ± 0.01

0.11 ± 0.01

0.13 ± 0.07

0.28 ± 0.11

Cleft width (nm)

Lateral

17.22 ± 1.50

15.52 ± 0.39

14.11 ± 0.69

17.24 ± 2.21

Central

30.22 ± 1.42

31.32 ± 1.81

16.47 ± 1.85

19.05 ± 2.72

Synaptic vesicles

Total number

4846+ / 4861-

5032+ / 5233-

740 ± 285

561.00 ± 108.00

811.47 ± 272.25

337± 23

555 ± 48

1820.64 ± 980.34

1518.52 ± 303.18

Diameter (nm)

29.85 ± 4.63

33.75 ± 4.55

~35

~35

19.80 ± 5.63

36.69 ± 1.71

Volume (µm3)

0.01 ± 0.00

0.02 ± 0.02

0.01 ± 0.01

Pool size of synaptic vesicles

Putative RRP at p10 nm

1.97 ± 2.57

3.89 ± 3.35

20.20 ± 18.58

5.42 ± 4.09

Putative RRP at p20 nm

6.30 ± 6.40

11.55 ± 4.16

48.59 ± 39.02

15.21 ± 9.09

Putative RP 60–200 nm

130.16 ± 20.79

162.83 ± 56.37

382.10 ± 248.23

181.86 ± 24.20

Putative resting pool ˃200nm

408.84 ± 100.04

599 ± 212.21

1251.82 ± 471.17

1264.05 ± 269.91

The various synaptic parameters were taken from *Bopp et al. 2017 values are taken from VGluT2-labeled boutons (+) or unlabeled VGluT2 boutons (−) density in µm2; **Rodriguez-Moreno et al. 2018; ***Hsu et al. 2017; Medalla and Luebke 2015 density per neuron; Rat Barrel Cortex: L4 and L5, Rollenhagen et al. 2015, 2018; Human TL: L4 and L5, Yakoubi et al. 2019a, b density in mm3.

Abbreviations: L, layer; TL, temporal lobe; M1, primary motor; S1, primary sensory; V1, primary visual; FC, frontal cortex; LPFC, lateral prefrontal cortex;

p (perimeter) 10 nm and p20 nm is the distance of synaptic vesicles from the active zone

Very special entities: Synaptic boutons in humans?

One of the major question in synaptic neuroscience is whether results obtained in experimental animals can be transferred one-to-one to the human brain. Research on the human brain was for a long-time restricted to postmortem brains. However, for fine-scale, high-resolution EM it turned out that tissue samples from postmortem brains are not suitable because the time window between the removal and fixation of the brain is far too long to guarantee an excellent preservation of the ultrastructure, a pre-requisite for EM investigations at the cellular and subcellular level. To overcome this problem access, non-epileptic tissue from epilepsy- or brain tumor surgery became the method of choice. Here, care was taken that the tissue samples were selected far away from the epileptic focusas monitored by magnetic resonance imaging and electrophysiology and may thus be regarded as non-affected (non-epileptic) as also demonstrated by other studies using the same experimental approach (Alonso-Nanclares et al. 2008; Navarrete et al. 2013; Mohan et al. 2015; Molnar et al. 2016; Seeman et al. 2018). After its removal, tissue sample can be either immediately immersion-fixed or even prepared for acute brain slice preparations. Meanwhile several studies have studied structural (for example: Alonso-Nanclares et al. 2008; Blazquez-Llorca et al. 2013; Morales et al. 2014; Liu and Schumann 2014, Yakoubi et al. 2019a, b) and functional aspects (for example: Holderith et al. 2016; Molnar et al. 2016; Seeman et al. 2018) of synaptic transmission and plasticity in humans. However, coherent and comprehensive studies about the synaptic organization of the human brain, in particular quantitative 3D-models of synaptic boutons in humans are still very rare (but see Yakoubi et al. 2019a, b).

Using non-affected neocortical access tissue taken from epilepsy surgery, we have started to study the layer-specific synaptic organization of the temporal lobe neocortex (TLN), a typical example of a six-layered granular associational neocortex (Fig. 2–4B, C). The growing interest in the TLN is motivated by its importance in high-order brain functions as audition, vision, memory, language processing, and various multimodal associations. Moreover, the TLN is also involved in several neurological diseases most importantly as the area of origin and onset of TL epilepsy (TLE). TLE is the most common form of refractory epilepsy characterized by recurrent, unprovoked focal seizures that may, with progressing disease, also spread to other areas of the brain. Taken together, the TLN represents an important region in the normal and pathologically altered brain in humans.

So far, the synaptic organization of layer 4, the receiving input layer of signals from the sensory periphery thus representing the first station of intracortical information processing and layer 5 the major output layer was quantitatively analyzed in the TLN (Yakoubi et al. 2019a, b). The final goal of our investigations is to describe the synaptic organization of a cortical column, the elementary building block of the neocortex also in humans, exemplified for the TLN.

Synaptic boutons in the human TLN have an average size of ~2.5 to 6 µm2 and are, beside similarities, strikingly different in some structural parameters from their counterparts in experimental animals (Table 1). Like in rodents and non-human primates so-called en passant (Fig. 2B; 4A) and endterminal synaptic boutons (Fig. 2C) contact either dendritic shafts (Figs. 2D; 3A, 4B), but the vast majority (~90 %) of excitatory synaptic boutons was established on dendritic spines of different types including stubby (Figs. 2D, 3A, B), mushroom (Figs. 3C, 4A, C), filopodial and elongated spines which is different to various animal species. Secondly, the majority of spines (~90 %) contained a so-called spine apparatus (Figs. 2D, 3A, B), a derivate of the endoplasmic reticulum, responsible for spine motility and stabilization of the synaptic complex during single or repetitive high-frequency stimulation. Thus, it was hypothesized that spines containing a spine apparatus partially contribute in modulating short-term synaptic plasticity (for example Holtmaat et al. 2006; for review see Knott and Holtmaat 2008). Interestingly, so-called dendro-dendritic synapses, regarded as a feature of the developmental brain, occur more frequently in the human TLN when compared to the neocortex in experimental animals. In addition, so-called clathrin-coated pits were frequently observed in synaptic boutons, some of which are located near the active zone (Fig. 2C). Clathrin-coated vesicles selectively sort cargo at the cell membrane, trans-Golgi network, and endosomal compartments for multiple membrane traffic pathways, for example exo- and endocytosis. A subpopulation is used in synaptic vesicle formation at the active zone.

Finally, also astrocytes receive direct synaptic input (Fig. 3D), although infrequently, supporting their involvement in synaptic transmission and plasticity (Min and Nevian 2012). Astrocytes have long been thought to act as nutrition suppliers and providing a stabilizing corset for neurons in the brain. However, it is now well established that astrocytes also play an important role in synaptic function, acting not only as physical barriers to glutamate diffusion, but also mediate transmitter uptake by glutamate transporters (Min and Nevian 2012; for review see Allen 2014; Dallerac et al. 2018). A striking common feature in both the human and the animal neocortex is the tight ensheathment of synaptic complexes with astrocytic processes forming the ‘tripartite’ synaptic complex (Fig. 3A, C), in contrast to MFBs and calyx of Held synapses, where astrocytic processes were never located close to individual active zones (Rollenhagen et al. 2007; Müller et al. 2009). This may explain the occurrence of glutamate spillover, synaptic cross talk and the switch from asynchronous to synchronous release upon repetitive stimulation as shown for the MFB (Hallermann et al. 2003) and Calyx of Held synapses (reviewed by von Gersdorff and Borst 2002). Astrocytes can actively take-up excessive or ‘spilled’ neurotransmitter when close to the synaptic cleft; hence they modulate the temporal and spatial neurotransmitter concentration thus controlling the induction, maintenance and termination of synaptic transmission but also modulate short-term synaptic plasticity in the neocortex.

Fig. 2: Synaptic organization in different layers of the human TLN
A, Low power TEM micrograph of the neuropil in layer 2/3. For better visualization, some structures are highlighted in different colors. Transparent magenta: a pyramidal cell (pyr) and an astrocyte (ast) close to another adjacent pyramidal cell. The nuclei are given in transparent blue. Numerous dendritic profiles of different shape and size in the neuropil are highlighted in transparent yellow and synaptic boutons in green. Mitochondria in all structures are given in transparent blue. Note the axon initial segment (ais) originating at the base of one pyramidal cell soma. The ais is innervated by several synaptic boutons (marked by asterisks). Scale bar 5 µm.

			B, En passant axon (ax) giving rise to a synaptic bouton (sb, transparent yellow) innervating a dendritic spine (sp, transparent blue) in layer 2/3. Active zones are outlined in red. Scale bar 2 µm.
			C, Two endterminal boutons (sb1, sb2) establishing synaptic contacts with two neighboring spines (sp1, sp2) in layer 4. Note the different shape and size of the boutons and the active zones (red contours) in both synaptic complexes that nearly covers the entire pre- and postsynaptic apposition zone. A so-called coated pit (asterisk) is located close to the active zone. Scale bar 0.2 µm.
			D, Representative example of a large stubby spine (sp) emerging from a dendrite (de) in layer 1 innervated by a large synaptic bouton (sb) with two active zones (marked by arrowheads) full of synaptic vesicles. Note also the prominent spine apparatus (framed area). Scale bar 0.5 µm.
			E, Large mushroom spine (sp) with a long spine neck (spn) emerging from a small dendrite (de) in layer 1. Note the disk-like shape of the prominent spine apparatus (red contours) that covers nearly the entire volume of the spine head. The spine head is innervated by three synaptic boutons (sb) and a fourth establishes a synaptic contact at the base of the spine neck. Scale bar 0.5 µm.
Fig. 2:

Synaptic organization in different layers of the human TLN

A, Low power TEM micrograph of the neuropil in layer 2/3. For better visualization, some structures are highlighted in different colors. Transparent magenta: a pyramidal cell (pyr) and an astrocyte (ast) close to another adjacent pyramidal cell. The nuclei are given in transparent blue. Numerous dendritic profiles of different shape and size in the neuropil are highlighted in transparent yellow and synaptic boutons in green. Mitochondria in all structures are given in transparent blue. Note the axon initial segment (ais) originating at the base of one pyramidal cell soma. The ais is innervated by several synaptic boutons (marked by asterisks). Scale bar 5 µm.

B, En passant axon (ax) giving rise to a synaptic bouton (sb, transparent yellow) innervating a dendritic spine (sp, transparent blue) in layer 2/3. Active zones are outlined in red. Scale bar 2 µm.

C, Two endterminal boutons (sb1, sb2) establishing synaptic contacts with two neighboring spines (sp1, sp2) in layer 4. Note the different shape and size of the boutons and the active zones (red contours) in both synaptic complexes that nearly covers the entire pre- and postsynaptic apposition zone. A so-called coated pit (asterisk) is located close to the active zone. Scale bar 0.2 µm.

D, Representative example of a large stubby spine (sp) emerging from a dendrite (de) in layer 1 innervated by a large synaptic bouton (sb) with two active zones (marked by arrowheads) full of synaptic vesicles. Note also the prominent spine apparatus (framed area). Scale bar 0.5 µm.

E, Large mushroom spine (sp) with a long spine neck (spn) emerging from a small dendrite (de) in layer 1. Note the disk-like shape of the prominent spine apparatus (red contours) that covers nearly the entire volume of the spine head. The spine head is innervated by three synaptic boutons (sb) and a fourth establishes a synaptic contact at the base of the spine neck. Scale bar 0.5 µm.

The most striking difference between synaptic boutons in the human, non-human primate and rodent neocortex is, however, the shape and size of the active zones and that of the three functionally defined pools of synaptic vesicles, namely the RRP, RP and resting pool. Although small in size synaptic excitatory synaptic boutons in layer 4 and layer 5 of the TLN contain active zones that were on average 2-fold larger in size (~0.2 – 0.25 µm2 in surface area) when compared with their counterparts of comparable size in other brain regions in rodents or non-human primates (Table 1), or even much larger CNS synapses such as the cerebellar and hippocampal mossy fiber bouton and the Calyx of Held endterminal. In numerous synaptic boutons in the human TLN, the active zones covered most of the pre- and postsynaptic apposition zone (Figs. 2 C, 3A, B) hence enlarging the presynaptic ‘docking’ zone for synaptic vesicles. In addition, more synaptic boutons contain not only a single but up to three active zones (Figs. 2B, 3A-C). Numerous of the active zones were perforated at the pre-, post- or both synaptic densities. Even more striking these boutons containing a total pool of synaptic vesicles (average 1500–1800 synaptic vesicles) that was 2–3-fold larger to that reported in the rodent and non-human primate neocortex suggesting also comparable large RRPs (Fig. 3F), RPs and resting pools. Indeed, the RRPs are by 3–5-fold, the RPs by 2-fold and the resting pools by 2-fold larger than in rodent and non-human primate neocortex. It has been recently shown that the size of the RRP dynamically regulates multivesicular release in mice (Vaden et al. 2019). Thus these large pools suggest reliable synaptic transmission even at high-frequency stimulation; hence a rapid depletion of the RRP and RP could be prevented by replenishment of synaptic vesicles from a large resting pool. It has to be noted though that like in non-human primates and rodents, a huge variability exists in the structural composition between individual synaptic boutons in humans, in particular the size of RRP, RP and resting pool that may partially contribute in modulating synaptic plasticity.

Fig. 3: Innervation pattern of synaptic boutons in different layers of the human TLN
A, Two opposite synaptic complexes (sb1-sp) and (sb2-de) in layer 1 one of which (sb1) establishes a glutamatergic synapse with a stubby spine (sp) identified by the shape and appearance of the active zone (arrowheads) and the size and more roundish shape of the synaptic vesicles. Sb2 is also glutamatergic as identified by the appearance of the active zone (arrowheads) but directly terminating on the dendritic shaft. Note the large astrocytic finger (ast) close to the dendrite and synaptic boutons and the spine apparatus in the stubby spine. Scale Bar 0.5 µm

			B, A large stubby spine (sp) in layer 2/3 innervated by a synaptic bouton (sb) with two separated active zones (arrowheads). The spine apparatus is marked by an asterisk. Scale bar 0.5 µm
			C, Typical example of a large mushroom spine in layer 2/3 with a thick spine head (sph) and a smaller but thick spine neck (spn) receiving input by a large synaptic bouton (sb). Note the two active zones (arrowheads) one directly on top of the spine head and the other located aside. The relatively large active zone (arrowheads) covers the entire pre- and postsynaptic apposition zone. de: dendrite; ast: astrocytic profile. Scale bar 0.5 µm.
			D, Astrocytic process (ast) identified by its opaque appearance and the vesicles containing gliotransmitter contacted by a synaptic bouton (sb) in layer 5. This type of contact is rarely found. Scale bar 0.5 µm.
			E, Dendro-dendritic synapse (de1, de2) where de1 serves as the presynaptic element identified by the cluster of synaptic vesicles at the active zone (arrowheads) in layer 6. In addition, de2 receives a synaptic bouton (sb) with a non-perforated active zone marked by arrowheads. Scale bar 0.25 µm.
			F, High-magnification of a large non-perforated active zone in layer 6 showing three ‘docked’ vesicles (highlighted in transparent green) among the population of synaptic vesicles close to the presynaptic density contacting a dendritic spine at the postsynaptic (post) element. Scale bar 0.25 µm.
Fig. 3:

Innervation pattern of synaptic boutons in different layers of the human TLN

A, Two opposite synaptic complexes (sb1-sp) and (sb2-de) in layer 1 one of which (sb1) establishes a glutamatergic synapse with a stubby spine (sp) identified by the shape and appearance of the active zone (arrowheads) and the size and more roundish shape of the synaptic vesicles. Sb2 is also glutamatergic as identified by the appearance of the active zone (arrowheads) but directly terminating on the dendritic shaft. Note the large astrocytic finger (ast) close to the dendrite and synaptic boutons and the spine apparatus in the stubby spine. Scale Bar 0.5 µm

B, A large stubby spine (sp) in layer 2/3 innervated by a synaptic bouton (sb) with two separated active zones (arrowheads). The spine apparatus is marked by an asterisk. Scale bar 0.5 µm

C, Typical example of a large mushroom spine in layer 2/3 with a thick spine head (sph) and a smaller but thick spine neck (spn) receiving input by a large synaptic bouton (sb). Note the two active zones (arrowheads) one directly on top of the spine head and the other located aside. The relatively large active zone (arrowheads) covers the entire pre- and postsynaptic apposition zone. de: dendrite; ast: astrocytic profile. Scale bar 0.5 µm.

D, Astrocytic process (ast) identified by its opaque appearance and the vesicles containing gliotransmitter contacted by a synaptic bouton (sb) in layer 5. This type of contact is rarely found. Scale bar 0.5 µm.

E, Dendro-dendritic synapse (de1, de2) where de1 serves as the presynaptic element identified by the cluster of synaptic vesicles at the active zone (arrowheads) in layer 6. In addition, de2 receives a synaptic bouton (sb) with a non-perforated active zone marked by arrowheads. Scale bar 0.25 µm.

F, High-magnification of a large non-perforated active zone in layer 6 showing three ‘docked’ vesicles (highlighted in transparent green) among the population of synaptic vesicles close to the presynaptic density contacting a dendritic spine at the postsynaptic (post) element. Scale bar 0.25 µm.

Fig. 4: 3D-volume reconstructions of synaptic boutons and their target structures and electron microscopic tomography in the rodent and human temporal lobe neocortex
A, En passant axon (ax, transparent gold) followed over long-distance in consecutive electron micrographs establishing a synaptic contact on a dendritic spine (sp) of a postsynaptic dendritic segment (de, blue) in rodent layer 5. The active zone is marked by arrowheads. Note the association of mitochondria (white) with the pool of synaptic vesicles (green dots). Note the presence of synaptic (green dots) vesicles and dense-core vesicles (magenta dots) within the en passant axon (framed area). Scale bar 1 µm.

			B, Two synaptic boutons terminating opposite to each other onto a small caliber dendrite (de) in layer 5. In one bouton, the envelope of the terminal is omitted to better visualize the distribution of synaptic (green dots) and dense-core (magenta) vesicles. The mitochondrion is given in white and the two active zones are marked by arrowheads. Scale bar 1 µm.
			C, Synaptic bouton (sb) terminating on a dendritic spine (sp) as visualized with electron microscopic tomography. A coated pit fused with the bouton membrane is shown in the framed area. Inset: High-power magnification of the active zone (arrowheads) with two ‘docked’ vesicles highlighted in transparent green. Pre: presynaptic; post: postsynaptic. Scale bar 0.25 µm.
Fig. 4:

3D-volume reconstructions of synaptic boutons and their target structures and electron microscopic tomography in the rodent and human temporal lobe neocortex

A, En passant axon (ax, transparent gold) followed over long-distance in consecutive electron micrographs establishing a synaptic contact on a dendritic spine (sp) of a postsynaptic dendritic segment (de, blue) in rodent layer 5. The active zone is marked by arrowheads. Note the association of mitochondria (white) with the pool of synaptic vesicles (green dots). Note the presence of synaptic (green dots) vesicles and dense-core vesicles (magenta dots) within the en passant axon (framed area). Scale bar 1 µm.

B, Two synaptic boutons terminating opposite to each other onto a small caliber dendrite (de) in layer 5. In one bouton, the envelope of the terminal is omitted to better visualize the distribution of synaptic (green dots) and dense-core (magenta) vesicles. The mitochondrion is given in white and the two active zones are marked by arrowheads. Scale bar 1 µm.

C, Synaptic bouton (sb) terminating on a dendritic spine (sp) as visualized with electron microscopic tomography. A coated pit fused with the bouton membrane is shown in the framed area. Inset: High-power magnification of the active zone (arrowheads) with two ‘docked’ vesicles highlighted in transparent green. Pre: presynaptic; post: postsynaptic. Scale bar 0.25 µm.

Taken together, the structural composition of both the presynaptic terminal and the spine as the main target structure suggests high synaptic efficacy and reliability of synaptic transmission but also in the induction, regulation and termination of short-term plasticity at synaptic boutons in the human brain.

This structural heterogeneity was confirmed by a recent structural/functional investigation about synaptic connections in the human TLN using paired recordings (Seeman et al. 2018). This study demonstrated that synaptic connections in the TLN are indeed highly reliable and strong as indicated by large excitatory postsynaptic potential (EPSP) amplitudes when compared to mouse neocortex, but also show layer-specific differences and in modulating short-term plasticity.

Perspectives

In summary, excitatory synaptic boutons in the brain represent ‘unique entities’ in rodents, non-human primates and even more in humans. Their individual composition with marked structural differences strongly suggest that they are perfectly adapted to the ‘job’ they have to fulfill in different neural networks of the brain in which they are embedded. However, there are still a lot of questions remaining that have to be addressed in the future. For example, and most importantly how and when synapses are generated during the development of the human brain and do they undergo the same selective layer-specific pruning or elimination as shown in experimental animals? How and when do they undergo severe structural and functional changes during neurological disorders like schizophrenia, autism and neurodegenerative diseases like Morbus Alzheimer and Morbus Parkinson? How age-, strain-, sex, left/right-, and subregion-specific differences would influence their structural composition? How the three functionally defined pools of synaptic vesicles become differentially recruited during high-frequency brain activity, during different biological rhythms or behavior still remains rather unclear. At the molecular level, how are the numerous pre- and postsynaptic proteins, neurotransmitters and their subunits involved in the induction, maintenance and termination of synaptic transmission and plasticity arranged at the active zone? What is about their density and possible co-localization at individual synaptic complexes? To date we are still far away in our understanding of these fascinating structures.

About the authors

Joachim H. R. Lübke

Born 11.04.1956 in Recklinghausen, Germany; 1974–1977: Education as a biological technician; 1981–1982: ‘Begabten-Abitur’ at the Abend-gymnasium Göttingen; 1982–1987: Study of Biology, Georg-August Universität Göttingen; 1987–1991: Diploma and doctoral thesis, Max-Planck-Institut für Biophysikalische Chemie, Abt. Neurobiologie (Prof. Dr. O.-D. Creutzfeldt); 1991–1993: Postdoctoral Fellow (Scolarship of the Royal Society of Science), Dept. of Human Anatomy (Prof. Dr. Ray Guillery, University of Oxford; 1993–1995: ‘Von Helmholtz Stipendiat’ of the BMBF, Anatomisches Institut (Prof. Dr. Michael Frotscher), Albert-Ludwigs Universität Freiburg; 1996: Wolfgang-Bargmann Preis, Anatomische Gesellschaft; 1996: Fachanatom, Anatomische Gesellschaft; 1999: Habilitation in Anatomy and Neuroanatomy, Albert-Ludwigs Universität Freiburg; 2000: Appointment as a C2 Lecturer, Albert-Ludwigs Universität Freiburg; 2003: Tenured-track appointment and Group leader position ‘Cellular Neurobiology’, Institute of Medicine (Prof. Dr. Karl Zilles, Research Centre Jülich GmbH; 2005: APL-Professor in Anatomy and Neuroanatomy, Heinrich-Heine Universität Düsseldorf; Since 2008: Professor W2, Medical Faculty, University Hospital/RWTH Aachen. Current Position: W2-Proffesor and group leader ‘Structure of Synapses’, Institute of Neuroscience and Medicine INM-10, Research Centre Jülich GmbH.

Astrid Rollenhagen

Born 06.11.1959 in Bielefeld, Germany; 1976–1978: Education as a doctor’s assistant; 1980–1982 ‘Abitur’ at the Westfalen Kolleg Bielefeld; 1982–1987: Study of Biology, Universität Bielefeld; 1987–1996: Diploma and doctoral thesis, Universität Bielefeld, Lehrstuhl Verhaltensforschung, AG Prof. Dr. H.-J. Bischof; 1996–1997: Scientific Assistant, Universität Bielefeld, Lehrstuhl Verhaltensforschung, AG Prof. Dr. H.-J. Bischof; 1998–2002: Universität Hamburg, Center for Molecular Neurobiology Hamburg, Abt. Prof. Dr. M. Schachner; 2002–2004: Albert-Ludwigs-Universität Freiburg, Dept. of Anatomy (Prof. Dr. M. Frotscher, AG PD Dr. J.H.R. Lübke); Since 2004: Senior Post-Doc, Group: ‘Structure of Synapses’ (Prof. Dr. J.H.R. Lübke), INM-10, Research Centre Jülich GmbH.

Acknowledgements

The authors are very grateful to all past and present members of the Group ‘Structure of Synapses’ at the INM-10 and INM-2, Research Centre Jülich GmbH for their individual contributions that made this review possible. We would also like to thank our collaborator Dr. Mike Hasenberg and his team at the IMCES Electron Microscopy Unit (EMU), Medical Research Centre, University Hospital Essen for their constant support with FIB-SEM and EM tomography. Finally, many thanks to Dr. Dorothea Miller and Priv. Doz. Dr. Marec von Lehe, Department of Neurosurgery, Knappschaftskrankenhaus Bochum and Brandenburg Medical School, Ruppiner Clinics, Neuruppin for providing the human tissue samples. Finally, the funding by the Helmholtz-Society and Deutsche Forschungsgemeinschaft is very much acknowledged.

Abbreviations in alphabetic order

Ca2+

calcium

CA3

hippocampal subregion CA3

CNS

central nervous system

CRYO-CLEM

CRYO-correlative light- and electron microscopy

3D

three-dimensional

dSTORM

stochastic optical reconstruction microscopy

EM

electron microscopy

EPSCs

excitatory postsynaptic currents

EPSPs

excitatory postsynaptic potentials

FIB-SEM

focused ion beam scanning electron microscopy

RRP

readily releasable pool of synaptic vesicles

RP

recycling pool of synaptic vesicles

STED

stimulated emission depletion microscopy

TEM

transmission electron microscopy

TLE

temporal lobe epilepsy

TLN

temporal lobe neocortex

References

Alle, H., and Geiger, J.R. (2006). Combined analog and action potential coding in hippocampal mossy fibers. Science 311, 1290–1293.10.1126/science.1119055Search in Google Scholar PubMed

Allen, N.J. (2014). Astrocyte regulation of synaptic behavior. Ann Rev Cell Dev Biol 30, 439–463.10.1146/annurev-cellbio-100913-013053Search in Google Scholar PubMed

Alonso-Nanclares, L., Gonzalez-Soriano, J., Rodriguez, J.R., and DeFelipe, J. (2008). Gender differences in human cortical synaptic density. Proc Natl Acad Sci USA 105, 14615–14619.10.1073/pnas.0803652105Search in Google Scholar PubMed PubMed Central

Anderson, J.C., and Martin, K.A.C. (2006). Synaptic connections from cortical areas V1 and V2 in macaque monkey. J Comp Neurol 495, 709–721.10.1002/cne.20914Search in Google Scholar PubMed

Anderson, J.C., and Martin, K.A.C. (2009). The synaptic connections between cortical area V4 to V2 in macaque monkey. J Neurosci 29, 11283–11293.10.1523/JNEUROSCI.5757-08.2009Search in Google Scholar PubMed PubMed Central

Bertram, R., Smith, G.D., and Sherman, A. (1999). Modeling study of the effects of overlapping Ca2+ microdomains on neurotransmitter release. Biophys J 76, 735–750.10.1016/S0006-3495(99)77240-1Search in Google Scholar PubMed PubMed Central

Biro, A.A., Holderith, N.B., and Nusser, Z. (2005). Quantal size is independent of the release probability at hippocampal excitatory synapses. J Neurosci 25, 223–232.10.1523/JNEUROSCI.3688-04.2005Search in Google Scholar PubMed PubMed Central

Bischofberger, J., and Jonas, P. (2002). TwoB or not twoB: differential transmission at glutamatergic mossy fiber-interneuron synapses in the hippocampus. Trends Neurosci 25, 600–603.10.1016/S0166-2236(02)02259-2Search in Google Scholar PubMed

Blazquez-Llorca, L., Merchan-Perez, A., Rodriguez, J.R., Gascon, J., and DeFelipe, J. (2013). FIB/SEM technology and Alzheimer’s disease: three-dimensional analysis of human cortical synapses. J Alzheimers Dis 34, 995–101310.3233/JAD-122038Search in Google Scholar PubMed

Bollmann, J.H., Sakmann, B., and Borst, J.G. (2000). Calcium sensitivity of glutamate release in the calyx-type terminal. Science 289, 953–957.10.1126/science.289.5481.953Search in Google Scholar PubMed

Bopp, R., Holler-Rickauer, S., Martin, K.A., and Schuhknecht, G.F. (2017). An Ultrastructural Study of the Thalamic Input to Layer 4 of Primary Motor and Primary Somatosensory Cortex in the Mouse. J Neurosci 37, 2435–2448.10.1523/JNEUROSCI.2557-16.2017Search in Google Scholar PubMed PubMed Central

Borst, J.G., and Sakmann, B. (1996). Calcium influx and transmitter release in a fast CNS synapse. Nature 383, 431–434.10.1038/383431a0Search in Google Scholar PubMed

Borst, J.G., and Sakmann, B. (1998). Calcium current during a single action potential in a large presynaptic terminal of the rat brainstem. J Physiol 506, 143–157.10.1111/j.1469-7793.1998.143bx.xSearch in Google Scholar PubMed PubMed Central

Borst, J.G., and Sakmann, B. (1999). Depletion of calcium in the synaptic cleft of a calyx-type synapse in the rat brainstem. J Physiol 521, 123–133.10.1111/j.1469-7793.1999.00123.xSearch in Google Scholar PubMed PubMed Central

Chamberland, S., and Toth, K. (2016). Functionally heterogeneous synaptic vesicle pools support diverse synaptic signalling. J Physiol 594, 825–835.10.1113/JP270194Search in Google Scholar PubMed PubMed Central

Chicurel, M.E., and Harris, K.M. (1992). Three-dimensional analysis of the structure and composition of CA3 branched dendritic spines and their synaptic relationships with mossy fiber boutons in the rat hippocampus. J Comp Neurol 325, 169–182.10.1002/cne.903250204Search in Google Scholar PubMed

Dallerac, G., Zapata, J., and Rouach, N. (2018). Versatile control of synaptic circuits by astrocytes: where, when and how? Nat Rev Neurosci 19, 729–743.10.1038/s41583-018-0080-6Search in Google Scholar PubMed

Dufour, A., Rollenhagen, A., Sätzler, K., and Lübke, J.H.R. (2016). Development of Synaptic Boutons in Layer 4 of the Barrel Field of the Rat Somatosensory Cortex: A Quantitative Analysis. Cereb Cortex 26, 838–854.10.1093/cercor/bhv270Search in Google Scholar PubMed PubMed Central

Engel, D., and Jonas, P. (2005). Presynaptic action potential amplification by voltage-gated Na+ channels in hippocampal mossy fiber boutons. Neuron 45, 405–417.10.1016/j.neuron.2004.12.048Search in Google Scholar PubMed

Feldmeyer, D. (2012). Excitatory neuronal connectivity in the barrel cortex. Front Neuroanat 6, 24 doi: 10.3389/fnana.2012.00024. eCollection 2012.10.3389/fnana.2012.00024Search in Google Scholar PubMed PubMed Central

Feldmeyer, D., Brecht, M., Helmchen, F., Petersen, C.C., Poulet, J.F., Staiger, J.F., Luhmann, H.J., and Schwarz, C. (2013). Barrel cortex function. Prog Neurobiol 103, 3–27.10.1016/j.pneurobio.2012.11.002Search in Google Scholar PubMed

Freche, D., Pannasch, U., Rouach, N., and Holcman, D. (2011) Synapse geometry and receptor dynamics modulate synaptic strength. PLoS One 6, e25122. Doi 10.137/journal.pone.0025122.10.1371/journal.pone.0025122Search in Google Scholar PubMed PubMed Central

Freese, J.L., and Amaral, D.G. (2006) The synaptic organization of projections from the amygdala to visual cortical areas TE and V1 in the macaque monkey. J Comp Neurol 496, 655–667.10.1002/cne.20945Search in Google Scholar PubMed PubMed Central

Geiger, J.R., and Jonas, P. (2000). Dynamic control of presynaptic Ca(2+) inflow by fast-inactivating K(+) channels in hippocampal mossy fiber boutons. Neuron 28, 927–939.10.1016/S0896-6273(00)00164-1Search in Google Scholar

Hallermann, S., Pawlu, C., Jonas, P., and Heckmann, M. (2003). A large pool of releasable vesicles in a cortical glutamatergic synapse. Proc Natl Acad Sci USA 100, 8975–8980.10.1073/pnas.1432836100Search in Google Scholar PubMed PubMed Central

Harris, K.M., and Sultan, P. (1995). Variation in the number, location and size of synaptic vesicles provides an anatomical basis for the nonuniform probability of release at hippocampal CA1 synapses. Neuropharmacology 34, 1387–1395.10.1016/0028-3908(95)00142-SSearch in Google Scholar PubMed

Helmchen, F., Borst, J.G., and Sakmann, B. (1997). Calcium dynamics associated with a single action potential in a CNS presynaptic terminal. Biophys J 72, 1458–1471.10.1016/S0006-3495(97)78792-7Search in Google Scholar PubMed PubMed Central

Henze, D.A., Card, J.P., Barrionuevo, G., and Ben-Ari, Y. (1997). Large amplitude miniature excitatory postsynaptic currents in hippocampal CA3 pyramidal neurons are of mossy fiber origin. J Neurophysiol. 77, 1075–1086.10.1152/jn.1997.77.3.1075Search in Google Scholar PubMed

Holderith, N., Lorincz, A., Katona, G., Rozsa, B., Kulik, A., Watanabe, M., and Nusser, Z. (2016). Release probability of hippocampal glutamatergic terminals scales with the size of the active zone. Nat Neurosci 19, 172.10.1038/nn.3137Search in Google Scholar PubMed PubMed Central

Holtmaat, A., Wilbrecht, L., Knott, G.W., Welker, E., and Svoboda, K. (2006). Experience-dependent and cell-type-specific spine growth in the neocortex. Nature 441, 979–983.10.1038/nature04783Search in Google Scholar PubMed

Hsu, A., Luebke, J.I., and Medalla, M. (2017). Comparative ultrastructural features of excitatory synapses in the visual and frontal cortices of the adult mouse and monkey. J Comp Neurol 525,2175–2191.10.1002/cne.24196Search in Google Scholar PubMed PubMed Central

Imig, C., Min, S.W., Krinner, S., Arancillo, M., Rosenmund, C., Südhof, T.C., Rhee, J., Brose, N., and Cooper, B.H. (2014). The morphological and molecular nature of synaptic vesicle priming at presynaptic active zones. Neuron 84, 416–431.10.1016/j.neuron.2014.10.009Search in Google Scholar PubMed

Knott, G., and Holtmaat, A. (2008). Dendritic spine plasticity--current understanding from in vivo studies. Brain Res Rev. 58; 282–289. doi: 10.1016/j.brainresrev.01.002.10.1016/j.brainresrev.2008.01.002Search in Google Scholar PubMed

Liu, X.B., and Schumann, C.M. (2014). Optimization of electron microscopy for human brains with long-term fixation and fixed-frozen sections. Acta Neuropathol Commun 2, 42.10.1186/2051-5960-2-42Search in Google Scholar PubMed PubMed Central

Lübke, J., and Feldmeyer, D. (2007). Excitatory signal flow and connectivity in a cortical column: focus on barrel cortex. Brain Struct Funct 212, 3–17.10.1007/s00429-007-0144-2Search in Google Scholar PubMed

Marrone, D.F., LeBoutillier, J.C., and Petit, T.L. (2005). Ultrastructural correlates of vesicular docking in the rat dentate gyrus. Neurosci Lett 378, 92–97.10.1016/j.neulet.2004.12.019Search in Google Scholar PubMed

Medalla, M., and Luebke, J.I. (2015). Diversity of glutamatergic synaptic strength in lateral prefrontal versus primary visual cortices in the rhesus monkey. J Neurosci 35, 112–127. doi: 10.1523/JNEUROSCI.3426-14.2015.10.1523/JNEUROSCI.3426-14.2015Search in Google Scholar PubMed PubMed Central

Meinrenken, C.J., Borst, J.G., and Sakmann, B. (2002). Calcium secretion coupling at calyx of Held governed by nonuniform channel-vesicle topography. J Neurosci 22, 1648–1667.10.1523/JNEUROSCI.22-05-01648.2002Search in Google Scholar PubMed PubMed Central

Meinrenken, C.J., Borst, J.G., and Sakmann, B. (2003). Local routes revisited: the space and time dependence of the Ca2+ signal for phasic transmitter release at the rat calyx of Held. J Physiol 547, 665–689.10.1113/jphysiol.2002.032714Search in Google Scholar PubMed PubMed Central

Michanski, S., Smaluch, K., Steyer, A.M., Chakrabarti, R., Setz, C., Oestreicher, D., Fischer, C., Mobius, W., Moser, T., Vogl, C., et al. (2019). Mapping developmental maturation of inner hair cell ribbon synapses in the apical mouse cochlea. Proc Natl Acad Sci USA 116, 6415–6424.10.1073/pnas.1812029116Search in Google Scholar PubMed PubMed Central

Min,R.,and Nevian,T. (2012) Astrocyte signaling controls spike timing-dependent depression at neocortical synapses. Nat Neurosci 15, 746–753.10.1038/nn.3075Search in Google Scholar PubMed

Mohan, H., Verhoog, M.B., Doreswamy, K.K., Eyal, G., Aardse, R., Lodder, B.N., Goriounova, N.A., Asamoah, B., AB, B.B., Groot, C., et al. (2015). Dendritic and Axonal Architecture of Individual Pyramidal Neurons across Layers of Adult Human Neocortex. Cereb Cortex 25, 4839–4853.10.1093/cercor/bhv188Search in Google Scholar PubMed PubMed Central

Molnar, G., Rozsa, M., Baka, J., Holderith, N., Barzo, P., Nusser, Z., and Tamas, G. (2016). Human pyramidal to interneuron synapses are mediated by multi-vesicular release and multiple docked vesicles. eLIFE 5.10.7554/eLife.18167.008Search in Google Scholar

Morales, J., Benavides-Piccione, R., Dar, M., Fernaud, I., Rodríguez, A., Anton-Sanchez, L., Bielza, C., Larrañaga, P., DeFelipe, J., and Yuste, R. (2014). Random positions of dendritic spines in human cerebral cortex. J Neurosci. 34, 10078–10084. doi: 10.1523/JNEUROSCI.1085-14.2014.10.1523/JNEUROSCI.1085-14.2014Search in Google Scholar PubMed PubMed Central

Moser, T., Brandt, A., and Lysakowski, A. (2006). Hair cell ribbon synapses. Cell Tissue Res 326, 347–359.10.1007/s00441-006-0276-3Search in Google Scholar PubMed PubMed Central

Müller, J., Reyes-Haro, D., Pivneva, T., Nolte, C., Schaette, R., Lübke, J., and Kettenmann, H. (2009). The principal neurons of the medial nucleus of the trapezoid body and NG2(+) glial cells receive coordinated excitatory synaptic input. J Gen Physiol 134, 115–127.10.1085/jgp.200910194Search in Google Scholar PubMed PubMed Central

Navarrete, M., Perea, G., Maglio, L., Pastor, J., Garcia de Sola, R., and Araque, A. (2013). Astrocyte calcium signal and gliotransmission in human brain tissue. Cereb Cortex 23, 1240–1246.10.1093/cercor/bhs122Search in Google Scholar PubMed

Neher, E. (2015). Merits and Limitations of Vesicle Pool Models in View of Heterogeneous Populations of Synaptic Vesicles. Neuron 87, 1131–1142.10.1016/j.neuron.2015.08.038Search in Google Scholar PubMed

Nicol, M.J., and Walmsley, B. (2002). Ultrastructural basis of synaptic transmission between endbulbs of Held and bushy cells in the rat cochlear nucleus. J Physiol 539, 713–723.10.1113/jphysiol.2001.012972Search in Google Scholar PubMed PubMed Central

Plitzko, JM, Rigort, A, and Leis, A. (2009) Correlative cryo-light microscopy and cryo-electron tomography: from cellular territories to molecular landscapes. Curr Opin Biotechnol 20, 83–89.10.1016/j.copbio.2009.03.008Search in Google Scholar PubMed

Qi, G., Radnikow, G., and Feldmeyer, D. (2015). Electrophysiological and morphological characterization of neuronal microcircuits in acute brain slices using paired patch-clamp recordings. J Vis Exp, 52358.10.3791/52358Search in Google Scholar PubMed PubMed Central

Radnikow, G., and Feldmeyer, D. (2018). Layer- and Cell Type-Specific Modulation of Excitatory Neuronal Activity in the Neocortex. Front Neuroanat 12, 1.10.3389/fnana.2018.00001Search in Google Scholar PubMed PubMed Central

Rizzoli, S. O., and Betz, W.J. (2005). Synaptic vesicle pools. Nat Rev Neurosci. 6:57–69.10.1038/nrn1583Search in Google Scholar PubMed

Rodriguez-Moreno, J., Rollenhagen, A., Arlandis, J., Santuy, A., Merchan-Perez, A., DeFelipe, J., Lübke, J.H.R., and Clasca, F. (2018). Quantitative 3D Ultrastructure of Thalamocortical Synapses from the “Lemniscal” Ventral Posteromedial Nucleus in Mouse Barrel Cortex. Cereb Cortex 28, 3159–3175.10.1093/cercor/bhx187Search in Google Scholar PubMed PubMed Central

Rollenhagen, A., Sätzler, K., Rodriguez, E.P., Jonas, P., Frotscher, M., and Lübke, J.H.R. (2007). Structural determinants of transmission at large hippocampal mossy fiber synapses. J Neurosci 27, 10434–10444.10.1523/JNEUROSCI.1946-07.2007Search in Google Scholar PubMed PubMed Central

Rollenhagen, A., Klook, K., Sätzler, K., Qi, G., Anstötz, M., Feldmeyer, D., and Lübke, J.H.R. (2015). Structural determinants underlying the high efficacy of synaptic transmission and plasticity at synaptic boutons in layer 4 of the adult rat ‘barrel cortex’. Brain Struct Funct 220, 3185–3209.10.1007/s00429-014-0850-5Search in Google Scholar PubMed

Rollenhagen, A., Ohana, O., Sätzler, K., Hilgetag, C.C., Kuhl, D., and Lübke, J.H.R. (2018). Structural Properties of Synaptic Transmission and Temporal Dynamics at Excitatory Layer 5B Synapses in the Adult Rat Somatosensory Cortex. Front Synaptic Neurosci 10, 24.10.3389/fnsyn.2018.00024Search in Google Scholar PubMed PubMed Central

Rowland, K.C., Irby, N.K., and Spirou, G.A. (2000). Specialized synapse-associated structures within the calyx of Held. J Neurosci 20, 9135–9144.10.1523/JNEUROSCI.20-24-09135.2000Search in Google Scholar PubMed PubMed Central

Sätzler, K., Söhl, L.F., Bollmann, J.H., Borst, J.G., Frotscher, M., Sakmann, B., and Lübke, J.H.R. (2002). Three-dimensional reconstruction of a calyx of Held and its postsynaptic principal neuron in the medial nucleus of the trapezoid body. J Neurosci 22, 10567–10579.10.1523/JNEUROSCI.22-24-10567.2002Search in Google Scholar PubMed PubMed Central

Saviane, C., and Silver, R.A. (2006). Fast vesicle reloading and a large pool sustain high bandwidth transmission at a central synapse. Nature 439, 983–987.10.1038/nature04509Search in Google Scholar PubMed

Schikorski, T., and Stevens, C.F. (1997). Quantitative ultrastructural analysis of hippocampal excitatory synapses. J Neurosci 17, 5858–5867.10.1523/JNEUROSCI.17-15-05858.1997Search in Google Scholar PubMed PubMed Central

Schikorski, T., and Stevens, C.F. (1999). Quantitative fine-structural analysis of olfactory cortical synapses. Proc Natl Acad Sci USA 96, 4107–4112.10.1073/pnas.96.7.4107Search in Google Scholar PubMed PubMed Central

Schikorski, T., and Stevens C.F. (2001). Morphological correlates of functionally defined synaptic vesicle populations. Nat Neurosci. 4, 391–395.10.1038/86042Search in Google Scholar PubMed

Schneggenburger, R., Meyer, A.C., and Neher, E. (1999). Released fraction and total size of a pool of immediately available transmitter quanta at a calyx synapse. Neuron 23, 399–409.10.1016/S0896-6273(00)80789-8Search in Google Scholar PubMed

Schneggenburger, R., and Neher, E. (2000). Intracellular calcium dependence of transmitter release rates at a fast central synapse. Nature 406, 889–893.10.1038/35022702Search in Google Scholar PubMed

Schneggenburger, R., Sakaba, T., and Neher, E. (2002). Vesicle pools and short-term synaptic depression: lessons from a large synapse. Trends Neurosci 25, 206–212.10.1016/S0166-2236(02)02139-2Search in Google Scholar

Seeman, S.C., Campagnola, L., Davoudian, P.A., Hoggarth, A., Hage, T.A., Bosma-Moody, A., Baker, C.A., Lee, J.H., Mihalas, S., Teeter, C., et al. (2018). Sparse recurrent excitatory connectivity in the microcircuit of the adult mouse and human cortex. eLIFE 7.10.7554/eLife.37349Search in Google Scholar PubMed PubMed Central

Sikora, M.A., Gottesman, J., and Miller, R.F. (2005). A computational model of the ribbon synapse. J Neurosci Methods 145, 47–61.10.1016/j.jneumeth.2004.11.023Search in Google Scholar PubMed

Silver, R.A., Lübke, J., Sakmann, B., and Feldmeyer, D. (2003). High-probability uniquantal transmission at excitatory synapses in barrel cortex. Science 302, 1981–1984.10.1126/science.1087160Search in Google Scholar PubMed

Sorra, K.E., and Harris, K.M. (1993). Occurrence and three-dimensional structure of multiple synapses between individual radiatum axons and their target pyramidal cells in hippocampal area CA1. J Neurosci 13, 3736–3748.10.1523/JNEUROSCI.13-09-03736.1993Search in Google Scholar PubMed PubMed Central

Spacek, J., and Harris, K.M. (1998). Three-dimensional organization of cell adhesion junctions at synapses and dendritic spines in area CA1 of the rat hippocampus. J Comp Neurol 393, 58–68.10.1002/(SICI)1096-9861(19980330)393:1<58::AID-CNE6>3.0.CO;2-PSearch in Google Scholar

Studer, D., Zhao, S., Chai, X., Jonas, P., Graber, W., Nestel, S., and Frotscher, M. (2014). Capture of activity-induced ultrastructural changes at synapses by high-pressure freezing of brain tissue. Nat Protoc 9, 1480–1495.10.1038/nprot.2014.099Search in Google Scholar

Szabadics, J., Varga, C., Molnar, G., Olah, S., Barzo, P., and Tamas, G. (2006). Excitatory effect of GABAergic axo-axonic cells in cortical microcircuits. Science 311, 233–235.10.1126/science.1121325Search in Google Scholar

Takahashi, T., Forsythe, I.D., Tsujimoto, T., Barnes-Davies, M., and Onodera, K. (1996). Presynaptic calcium current modulation by a metabotropic glutamate receptor. Science 274, 594–597.10.1126/science.274.5287.594Search in Google Scholar

Vaden, J.H., Banumurthy, G., Gusarevich, E.S., Overstreet-Wadiche, L., and Wadiche, J.I. (2019). The readily-releasable pool dynamically regulates multivesicular release. eLIFE 8.10.7554/eLife.47434Search in Google Scholar

Verstreken, P., Ly, C.V., Venken, K.J., Koh, T.W., Zhou, Y., and Bellen, H.J. (2005). Synaptic mitochondria are critical for mobilization of reserve pool vesicles at Drosophila neuromuscular junctions. Neuron 47, 365–378.10.1016/j.neuron.2005.06.018Search in Google Scholar

Verstreken, P., Ohyama, T., and Bellen, H.J. (2008). FM 1–43 labeling of synaptic vesicle pools at the Drosophila neuromuscular junction. Methods Mol Biol.440, 349–69. doi: 10.1007/978-1-59745-178-926.10.1007/978-1-59745-178-9_26Search in Google Scholar

von Gersdorff, H., and Borst, J.G. (2002). Short-term plasticity at the calyx of Held. Nat Rev Neurosci 3, 53–64.10.1038/nrn705Search in Google Scholar

Wimmer, V.C., Horstmann, H., Groh, A., and Kuner, T. (2006). Donut-like topology of synaptic vesicles with a central cluster of mitochondria wrapped into membrane protrusions: a novel structure-function module of the adult calyx of Held. J Neurosci 26, 109–116.10.1523/JNEUROSCI.3268-05.2006Search in Google Scholar

Xu-Friedman, M.A., Harris, K.M., and Regehr, W.G. (2001). Threedimensional comparison of ultrastructural characteristics at depressing and facilitating synapses onto cerebellar Purkinje cells. J Neurosci 21, 6666–6672.10.1523/JNEUROSCI.21-17-06666.2001Search in Google Scholar

Xu-Friedman, M.A., and Regehr, W.G. (2003). Ultrastructural contributions to desensitization at cerebellar mossy fiber to granule cell synapses. J Neurosci 23, 2182–2192.10.1523/JNEUROSCI.23-06-02182.2003Search in Google Scholar PubMed PubMed Central

Yakoubi, R., Rollenhagen, A., von Lehe, M., Shao, Y., Sätzler, K., and Lübke, J.H.R. (2019a). Quantitative Three-Dimensional Reconstructions of Excitatory Synaptic Boutons in Layer 5 of the Adult Human Temporal Lobe Neocortex: A Fine-Scale Electron Microscopic Analysis. Cereb Cortex. 29, 2797–2814, doi: 10.1093/cercor/bhy146.10.1093/cercor/bhy146Search in Google Scholar PubMed

Yakoubi, R., Rollenhagen, A., von Lehe, M., Miller D., Walkenfort, B., Hasenberg, M., Sätzler, K., and Lübke, JHR. (2019b). Ultrastructural heterogeneity of human layer 4 excitatory synaptic boutons in the adult temporal lobe neocortex. Program No. 037.01. 2019 Neuroscience Meeting Planner. Chicago, IL: Society for Neuroscience, 2019. Online.10.1101/643924Search in Google Scholar

Yamada, W.M., and Zucker, R.S. (1992). Time course of transmitter release calculated from simulations of a calcium diffusion model. Biophys J 61, 671–682.10.1016/S0006-3495(92)81872-6Search in Google Scholar PubMed PubMed Central

For the online publication

Movie 1: Representative example of a z-stack of 100 consecutive images through layer 1b of the human temporal neocortex taken with a FIB-SEM. Note the rapid change in the organization of the neuropil. (Collaboration with Dr. Mike Hasenberg and his team at the IMCES Electron Microscopy Unit (EMU), Medical Research Centre, University Hospital Essen).

Movie 2: Representative example of a shaft synapse in layer 4 of the human temporal lobe neocortex as revealed by EM tomography. Note the occurrence of three mitochondria closely associated with the pool of synaptic vesicles in the presynaptic terminal and the rapid change in the shape and size of the active zone. Scale bar 0.2 µm. (Collaboration with Dr. Mike Hasenberg and his team at the IMCES Electron Microscopy Unit (EMU), Medical Research Centre, University Hospital Essen).

Published Online: 2019-12-05
Published in Print: 2020-02-25

© 2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 5.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/nf-2019-0015/html
Scroll to top button