Imaging Synapse Structure and Function
Imaging Synapse Structure and Function
STRUCTURE AND
FUNCTION
EDITED BY : George J. Augustine and Marc Fivaz
PUBLISHED IN : Frontiers in Synaptic Neuroscience
Frontiers Copyright Statement
About Frontiers
© Copyright 2007-2017 Frontiers
Media SA. All rights reserved. Frontiers is more than just an open-access publisher of scholarly articles: it is a pioneering
All content included on this site,
such as text, graphics, logos, button
approach to the world of academia, radically improving the way scholarly research
icons, images, video/audio clips, is managed. The grand vision of Frontiers is a world where all people have an equal
downloads, data compilations and
software, is the property of or is opportunity to seek, share and generate knowledge. Frontiers provides immediate and
licensed to Frontiers Media SA permanent online open access to all its publications, but this alone is not enough to
(“Frontiers”) or its licensees and/or
subcontractors. The copyright in the realize our grand goals.
text of individual articles is the property
of their respective authors, subject to
a license granted to Frontiers. Frontiers Journal Series
The compilation of articles constituting The Frontiers Journal Series is a multi-tier and interdisciplinary set of open-access, online
this e-book, wherever published,
as well as the compilation of all other journals, promising a paradigm shift from the current review, selection and dissemination
content on this site, is the exclusive
property of Frontiers. For the
processes in academic publishing. All Frontiers journals are driven by researchers for
conditions for downloading and researchers; therefore, they constitute a service to the scholarly community. At the same
copying of e-books from Frontiers’
website, please see the Terms for time, the Frontiers Journal Series operates on a revolutionary invention, the tiered publishing
Website Use. If purchasing Frontiers system, initially addressing specific communities of scholars, and gradually climbing up to
e-books from other websites
or sources, the conditions of the broader public understanding, thus serving the interests of the lay society, too.
website concerned apply.
Images and graphics not forming part
of user-contributed materials may
Dedication to Quality
not be downloaded or copied
Each Frontiers article is a landmark of the highest quality, thanks to genuinely collaborative
without permission.
Individual articles may be downloaded interactions between authors and review editors, who include some of the world’s best
and reproduced in accordance academicians. Research must be certified by peers before entering a stream of knowledge
with the principles of the CC-BY
licence subject to any copyright or that may eventually reach the public - and shape society; therefore, Frontiers only applies
other notices. They may not be
re-sold as an e-book.
the most rigorous and unbiased reviews.
As author or other contributor you Frontiers revolutionizes research publishing by freely delivering the most outstanding
grant a CC-BY licence to others to research, evaluated with no bias from both the academic and social point of view.
reproduce your articles, including any
graphics and third-party materials By applying the most advanced information technologies, Frontiers is catapulting scholarly
supplied by you, in accordance with
the Conditions for Website Use and
publishing into a new generation.
subject to any copyright notices which
you include in connection with your What are Frontiers Research Topics?
articles and materials.
All copyright, and all rights therein, Frontiers Research Topics are very popular trademarks of the Frontiers Journals Series:
are protected by national and
international copyright laws. they are collections of at least ten articles, all centered on a particular subject. With their
The above represents a summary unique mix of varied contributions from Original Research to Review Articles, Frontiers
only. For the full conditions see the
Conditions for Authors and the
Research Topics unify the most influential researchers, the latest key findings and historical
Conditions for Website Use. advances in a hot research area! Find out more on how to host your own Frontiers
ISSN 1664-8714
Research Topic or contribute to one as an author by contacting the Frontiers Editorial
ISBN 978-2-88945-175-3
DOI 10.3389/978-2-88945-175-3 Office: [email protected]
Frontiers in Synaptic Neuroscience 1 May 2017 | Imaging Synapse Structure and Function
IMAGING SYNAPSE STRUCTURE
AND FUNCTION
Topic Editors:
George J. Augustine, Lee Kong Chian School of Medicine, Nanyang Technological University,
Singapore
Marc Fivaz, Duke-NUS Medical School, Singapore
Frontiers in Synaptic Neuroscience 2 May 2017 | Imaging Synapse Structure and Function
permits observation of synapses and their molecular machinery at sub-diffraction resolution. At
the ultrastructural level, automated forms of electron microscopy, improvements in specimen
fixation methods, and recent efforts to correlate data from light and electron micrographs now
make the reconstruction of functional neural circuits a reality. Finally, the use of optogenetic
actuators, such as channelrhodopsins, allows precise temporal and spatial manipulation of neu-
ronal activity and is revealing profound insights into the organization of neural circuits and
their roles in behavior.
This research topic highlights recent advances in both light and electron microscopy, with a
specific focus on approaches that combine innovations from several different fields to obtain
novel information about synapse structure and function. We are confident that this collection
of articles - three original research papers, six reviews, one methods paper and one perspective
article - will enable neuroscientists to achieve the next generation of experiments aimed at
cracking the neural code.
Citation: Augustine, G. J., Fivaz, M., eds. (2017). Imaging Synapse Structure and Function.
Lausanne: Frontiers Media. doi: 10.3389/978-2-88945-175-3
Frontiers in Synaptic Neuroscience 3 May 2017 | Imaging Synapse Structure and Function
Table of Contents
Frontiers in Synaptic Neuroscience 4 May 2017 | Imaging Synapse Structure and Function
EDITORIAL
published: 15 December 2016
doi: 10.3389/fnsyn.2016.00036
These are the glory days for imaging synapse structure and function. Thanks to recent advances
in both optical and electron microscopy, it is now possible to image individual synapses with
unprecedented spatial and temporal resolution. The parallel development of a wide range
of genetically-encoded synaptic reporters enables all-optical recording of synaptic activity in
genetically-defined neuronal populations. These engineering breakthroughs allow neuroscientists
to interrogate the brain in ways that were inconceivable just a few years ago. The ability to image
synaptic structure and activity in large functional circuits is beginning to yield key insights into
how the brain stores, processes, and computes information. This research topic consists of eleven
articles (methods, primary research papers, and reviews) that provide an overview of the latest
developments in synapse imaging. Rather than attempting an exhaustive list of synaptic reporters
and microscopy techniques, our collection emphasizes approaches that merge technical advances
from diverse areas to extract a rich palette of novel information from individual synapses.
Watanabe presents a method that combines optogenetics and rapid freezing (Flash-and-Freeze)
to visualize the synaptic vesicle (SV) cycle at the ultrastructural level with millisecond resolution
Watanabe. This revolutionary approach revealed ultra-fast endocytosis of SVs at central synapses
and neuromuscular junctions (Watanabe et al., 2013a,b) and promises to uncover many new
kinetic aspects of synapse dynamics. Begemann and Galic review recent efforts to image neuronal
preparations with both light and electron microscopy, with a series of hybrid techniques referred
Edited and reviewed by: to as Correlative Light Electron Microscopy (CLEM). Jackson and Burrone describe the first
Per Jesper Sjöström, genetically-encoded fluorescent reporter (sypHy-RGECO) that enables concurrent monitoring
McGill University, Canada of calcium dynamics and SV fusion. sypHy-RGECO will undoubtedly be a powerful means of
*Correspondence: examining calcium triggering of SV exocytosis at the level of individual presynaptic boutons. Using
Marc Fivaz similar probes, Tang et al. show that the mental disease gene DISC1 (Disrupted-In-Schizophrenia-
[email protected] 1) accelerates SV exocytosis by facilitating calcium influx through N-type voltage-gated Ca2+
channels. Calcium transients at synapses are also shaped by both mobilization and sequestration
Received: 19 October 2016
of calcium by intracellular stores. Kwon et al. report on the latest advances in organelle-specific
Accepted: 22 November 2016
calcium sensors and review the contribution of the endoplasmic reticulum and mitochondria to
Published: 15 December 2016
calcium dynamics and synaptic transmission/plasticity.
Citation:
Until recently, one major impediment to imaging of synaptic activity has been our inability to
Augustine GJ and Fivaz M (2016)
Editorial: Imaging Synapse Structure
directly measure membrane potential with adequate signal/noise ratio. This is rapidly changing
and Function. with the recent improvement of a wide range of genetically-encoded voltage indicators (GEVIs)
Front. Synaptic Neurosci. 8:36. that now are capable of monitoring both single action potentials and even subthreshold
doi: 10.3389/fnsyn.2016.00036 synaptic potentials, both in vitro and in vivo Nakajima et al. Three papers describe recent
advances on the localization, dynamics and function of Overall, we hope that the fine collection of papers contained
postsynaptic receptors and scaffolds. Using a combination within this research topic highlights a useful synapse imaging
of single-molecule tracking (uPAINT) and photoactivated toolkit for the neuroscience community. The next big challenge
localization microscopy (PALM), Li and Blanpied assess in brain imaging will be to scale up these synaptic measurements
the diffusion properties of membrane proteins within the to large ensembles of neurons to comprehend how circuits
postsynaptic density (PSD). The same authors recently used compute. This will require synaptic reporters that operate in
localization microscopy to demonstrate the existence of a synaptically-relevant time scale (milliseconds), along with
transsynaptic nanocolumns that align the neurotransmitter improved genetic targeting strategies, further advances in
release machinery to postsynaptic receptors (Tang et al., 2016a). automated high-speed microscopy, and refined bioinformatics
Dosemeci et al. review a series of EM studies that reveal the tools for analysis of the resulting large datasets.
presence of a dense lamina—the “pallium”—just beneath the
core layer of the PSD, and discuss how translocation of signaling AUTHOR CONTRIBUTIONS
proteins and scaffolds in and out of the pallium may shape
activity-induced changes in dendritic spines. In keeping with the All authors listed, have made substantial, direct and intellectual
theme of postsynaptic signaling, Dore et al. discuss evidence for contribution to the work, and approved it for publication.
metabotropic functions of NMDARs, based on time-resolved
FRET and other imaging approaches. Finally, at the level of FUNDING
synaptic circuits, two reviews describe the use of genetically-
encoded synaptic labels to trace neural circuits in a variety of This work was supported by Singapore Ministry of Education
different model systems, ranging from C. elegans to mammals grants MOE2015-T1-001-069 and MOE2015-T2-2-095 to GA
(Hong and Park; Lee et al.). and grant MOE2013-T2-1-053 to MF.
REFERENCES Conflict of Interest Statement: The authors declare that the research was
conducted in the absence of any commercial or financial relationships that could
Tang, A. H., Chen, H., Li, T. P., Metzbower, S. R., Macgillavry, H. D., and Blanpied, be construed as a potential conflict of interest.
T. A. (2016a). A trans-synaptic nanocolumn aligns neurotransmitter release to
receptors. Nature 536, 210–214. doi: 10.1038/nature19058 Copyright © 2016 Augustine and Fivaz. This is an open-access article distributed
Watanabe, S., Liu, Q., Davis, M. W., Hollopeter, G., Thomas, N., Jorgensen, N. B., under the terms of the Creative Commons Attribution License (CC BY). The
et al. (2013a). Ultrafast endocytosis at Caenorhabditis elegans neuromuscular use, distribution or reproduction in other forums is permitted, provided the
junctions. Elife 2:e00723. doi: 10.7554/eLife.00723 original author(s) or licensor are credited and that the original publication in
Watanabe, S., Rost, B. R., Camacho-Perez, M., Davis, M. W., Söhl-Kielczynski, this journal is cited, in accordance with accepted academic practice. No use,
B., Rosenmund, C., et al. (2013b). Ultrafast endocytosis at mouse hippocampal distribution or reproduction is permitted which does not comply with these
synapses. Nature 504, 242–247. doi: 10.1038/nature12809 terms.
DISC1 is the prototypical example of a gene associated in large synapse populations. We show that DISC1 elevates
with several major psychiatric disorders. It was discovered synaptic Ca2+ signals and boosts SV exocytosis at glutamatergic
in a Scottish family at the site of a balanced chromosomal terminals. Our results further indicate that N-type voltage-gated
translocation that strongly segregates with major depression, Ca2+ channels (VGCCs) mediate the stimulatory effect of DISC1
schizophrenia and bipolar disorder (St Clair et al., 1990; on SV release. These findings identify a central role of DISC1
Millar et al., 2000, 2001). The high penetrance (∼70%) of in neurotransmitter release and provide new insights on the
this translocation for mental illness supports a causal link biological basis of synaptic dysfunction in major psychiatric
between this genetic lesion and major psychiatric conditions disorders.
(Chubb et al., 2008). DISC1 variants (haplotypes, single
nucleotide polymorphisms and copy number variations) MATERIALS AND METHODS
have since been independently associated with depression,
schizophrenia, bipolar disorders and autism spectrum DNA, shRNA Constructs, Lentiviruses and
disorders (Ekelund et al., 2001, 2004; Sachs et al., 2005;
Kilpinen et al., 2008). Thus, DISC1 is a major susceptibility
Antibodies
The pCAGGs vGlut1-pHluorin (vGpH) and pCAGGs
factor for mental illness and a relevant genetic entry point to
Synaptophysin-GCaMP3 (SyGC3) constructs were gifts from
identify core endophenotypes implicated in neuropsychiatric
R. Edwards (UCSF; Voglmaier et al., 2006) and S. Voglmaier
disorders.
(UCSF; Li et al., 2011). The pFUGW (Addgene #14883) shRNA-
The translocation breakpoint in this Scottish family is
expressing lentiviral vector was modified to express mCherry
located in the C-terminal portion of DISC1 and results in
(pFUmChW). The shRNA targeting sequences are as follows:
overall reduced expression of the full-length transcript and
(1) Scramble 50 GGAGCAGACGCTGAATTAC30 (Kamiya
protein (Millar et al., 2005), suggesting that haploinsufficiency
et al., 2005); (2) DISC1-E 50 GGCTACATGAGAAGCACAG30
is the main mechanism by which this chromosomal alteration
(exon 2; Duan et al., 2007); and (3) DISC1-A 50 GGAAGG
confers risk to disease. Alternatively, it has been proposed
GCTAGAGATGTTC30 (exon 9) designed with Block-it shRNA
that a C-terminal truncated form of DISC1 is expressed
from Invitrogen. pFUGW scramble shRNA was a gift from
from the translocated allele and may be pathogenic (Hikida
A. Sawa (Johns Hopkins). The DISC1 shRNAs constructs
et al., 2007; Pletnikov et al., 2008), although expression of the
were cloned by introducing double-stranded DNA oligos
truncated DISC1 protein in translocation carriers remains to
into lentiviral vector pll3.7 (Addgene #11795) using the
be demonstrated. Consistent with a disease mechanism based
HpaI and XhoI sites. DNA fragments containing the U6
on DISC1 loss-of-function, DISC1 expression is also attenuated
promoter and shRNAs were then PCR amplified and cloned
in human induced pluripotent stem (iPS) cells derived from
into pFUmChW using the PacI site. The human DISC1 gene
members of a family with a DISC1 frame-shift mutation that
L variant (NCBI Refseq NM018662.2) was PCR amplified
co-segregates with major psychiatric disorders (Wen et al.,
from pCMV6-XL5 DISC1-tGFP (Origene) and cloned into
2014).
pIRES2-DsRed-Express (Clontech) using NheI and SmaI sites.
The identification of a large DISC1 interactome consisting
All constructs were sequenced before use. Lentiviral particles
of proteins belonging to different ontologic families suggests
expressing pFUmChW shRNAs were prepared as described in
broad functions of DISC1 in nerve cells. Accordingly, DISC1
Tiscornia et al. (2006). The DNA constructs used for whole-
has been implicated in multiple aspects of neuronal and brain
cell patch clamp recording are as follows: Cav2.1 (generated
development, including neurogenesis (Clapcote et al., 2007;
in T. W. Soong’s lab), Cav2.2 (Addgene #26568), GFP-β2a
Shen et al., 2008; Mao et al., 2009; Singh et al., 2010; Lee
and α2 δ1 (kindly provided by T. Snutch, UBC). The rabbit
et al., 2011), neuronal migration (Kamiya et al., 2005; Duan
polyclonal Abs against Cav2.1 (#ACC-001) and Cav2.2 (#ACC-
et al., 2007; Kubo et al., 2010; Steinecke et al., 2012) and
002) were from Alomone Labs. The polyclonal Ab against
maturation (Duan et al., 2007; Shinoda et al., 2007; Niwa
the C-terminus of mouse DISC1 was previously described
et al., 2010). Even though DISC1 interacts with several signaling
(Kuroda et al., 2011). The polyclonal Ab against human DISC1
proteins known to regulate synaptic functions, relatively little
(ab59017) and monoclonal Ab against bassoon were from
is known about functions of DISC1 at the synapse. In
Abcam (ab82958).
particular, the impact of DISC1 on neurotransmitter release
remains largely unexplored, despite the fact that aberrant
dopamine and glutamate neurotransmission is a probable cause Mouse Lines, Primary Neuron Cultures and
of schizophrenia and other mood disorders (Howes et al., Transfection Protocols
2015). DISC1 (∆2–3) mice (C57BL/6JJmsSlc) have been described
Given the preferential expression of DISC1 in the before Kuroda et al. (2011). Heterozygous DISC1wt/∆2–3
hippocampus and the involvement of this brain structure mice were crossed to each other to obtain DISC1wt/wt and
in cognition and psychiatric disorders (Chubb et al., 2008), we DISC1∆2–3/∆2–3 lines from which hippocampal neurons were
set out to determine the impact of DISC1 on synaptic vesicle prepared from E17/E18 embryos, according to published
(SV) cycling in hippocampal neurons. We used two independent procedures using papain (Worthington) digestion (Huettner
genetic strategies to alter DISC1 expression—RNAi and a and Baughman, 1986). Primary rat hippocampal neurons
DISC1 KO mouse—and imaged SV cycling and Ca2+ dynamics were prepared from E18 embryos according to Kaech and
Banker (2006) and as previously described in Garcia-Alvarez (Sigma-Aldrich). Cells were then incubated with 5% goat
et al. (2015). For live-cell imaging and immunocytochemistry serum to block non-specific binding sites and stained with the
experiments, neurons were grown on poly-L-Lysine coated glass primary and Alexa Fluor-conjugated secondary antibodies (Life
coverslips, on top of a glial feeder according to the Banker Technologies). Coverslips were then mounted on glass slides
Protocol (Kaech and Banker, 2006). For biochemical analysis, and imaged with an inverted laser scanning confocal microscope
cells were seeded on poly-L-Lysine coated 6-well culture dishes. (LSM710, Zeiss) with a Plan-Apochromat 63× (NA = 1.40)
Unless otherwise stated, neurons were cultured for 14–16 days. objective.
The vGpH and SyGC3 constructs were electroporated in
freshly-dissociated neurons using the Nucleofector kit (Amaxa
Biosystems, Lonza). Lentiviral particles expressing shRNAs
Image Analysis
For automated analysis of vGpH responses, we wrote a
were added on DIV2, at a MOI (multiplicity of infection) of
Matlab script that segments responsive boutons based on
1–3. All animal procedures were approved by the SingHealth
the difference between peak vGpH intensity during the
Institutional Animal Care and Use Committee (IACUC) of
first stimulation and baseline intensity prior to stimulation
Singapore.
(∆F = F peak − F baseline ). An intensity threshold for ∆F was
selected to resolve individual boutons and exclude those with
Live-Cell Confocal Imaging and ∆F below 5%. The same threshold was used for all conditions
in one independent experiment (i.e., one neuron preparation
Immunofluorescence with control vs. DISC1-depleted conditions). This threshold
Time-lapse confocal imaging was performed on an inverted
value was minimally adjusted (less than 10% change) across
Eclipse TE2000-E microscope (Nikon, USA), mounted with
all independent experiments described in this article. Binarized
a spinning-disk confocal scan head (CSU-10; Yokogawa,
boutons were then slightly dilated (Figure 1B) to ensure that
Japan), and equipped with a temperature controlled (36.5◦ C)
vGpH fluorescence was captured in its totality even after minor
stage and an autofocusing system (PFS; Nikon). Images were
lateral movement or change in shape. Time series with minor
acquired with an Orca-Flash 4.0 CCD camera (Hamamatsu
x–y drifts (originating from the stage) were re-aligned using a
Photonics, Japan) controlled by MetaMorph 7.8.6 (Molecular
script previously described (Thevathasan et al., 2013). Segmented
Devices, CA, USA) at 0.5 Hz or 20 Hz. Samples were
boutons were then size gated, with gating parameters kept
imaged using a 60× (NA 1.4) objective in Tyrode’s buffer
constant across all experiments. vGpH fluorescence intensity
(150 mM NaCl, 2.5 mM KCl, 1 mM CaCl2 , 1 mM MgCl2 ,
was then extracted in each segmented bouton across the time
6 mM glucose, 25 mM HEPES, pH 7.4) supplemented with
series and divided by the signal after NH4 Cl to normalize
25 µM 6-cyano-7-nitroquinoxaline-2,3-dione/CNQX (Tocris
for vGpH expression (Figure 1C). This segmentation strategy
Bioscience) and 50 µM D,L-2-amino-5-phosphonovaleric
ensures that the same boutons are analyzed during the two
acid/AP5 (Tocris Bioscience). Coverslips were mounted in an
consecutive AP trains. vGpH traces with significant baseline
RC-21BRFS chamber (Warner Instruments, USA) equipped
drifts between the first and second stimulation or after Baf
with platinum wire electrodes. Field stimulation was induced
application were excluded. A similar segmentation approach
by a square pulse stimulator (Grass Technologies, USA)
was used to analyze SyGC3 Ca2+ signals. SV exocytic rates
and monitored by an oscilloscope (TDS210, Tektronix,
were obtained by measuring the slope of the vGpH rise
USA). Trains of action potentials (APs) were generated
during the second stimulation (in the presence of Baf). The
by applying 20 V pulses (1 ms duration) at 10 or 20 Hz.
first six time points (during stimulation) were used for linear
Our typical stimulation paradigm for vGpH measurements
regression analysis. To measure endocytic rates, the vGpH
involved two consecutive trains of 300 APs at 10 Hz,
trace during the first stimulation was subtracted from that
separated by ∼5 min to allow synapses to recover. vGpH
obtained during the second stimulation. The resulting trace
responses between the first and second stimulation were
is a measure of endocytosis (Figure 1D). Endocytic rates
highly reproducible (data not shown). For measuring SV
were measured by linear fitting of six time points chosen
exocytic rates, the second stimulation was preceded (30 s
after stimulation onset when endocytosis kicks in. Presynaptic
earlier) by the addition of Bafilomycin A1 (Baf; AG scientific,
localization, abundance and density of Cav2.1 or Cav2.2 were
USA) 0.5 µM. To normalize for total expression of vGpH
analyzed with a modified version of a Matlab script described
in each individual bouton, 50 mM NH4 Cl was added at
previously (Poon et al., 2014). All Matlab scripts are available
the end of the time series. For SyGC3 measurements, Ca2+
upon request.
signals were normalized by adding 10 µM or 50 µM (for
Figure 4G) ionomycin (Sigma-Aldrich) at the end of the
stimulation protocol. ω-agatoxin TK125 nM (Tocris Bioscience) Immunoblotting
and ω-conotoxin GVIA 125 nM (Alomone Labs) were Cultured primary neurons or HEK293T cells were washed
used for experiment involving inhibition of P/Q-type Ca2+ with ice-cold PBS and lysed with RIPA buffer (10 mM Tris-
channels and N-type Ca2+ channels activity, respectively. HCl pH = 7.2, 150 mM NaCl, 1% TritonX-100, 0.1% SDS,
For immunofluorescence studies, neurons grown on glass 5 mM EDTA, 0.25% Na-deoxycholate) supplemented with
coverslips were fixed in 4% paraformaldehyde with 4% Complete protease inhibitor and PhosphoStop phosphatase
sucrose in PBS and permeabilized with 100 ng/ml Digitonin inhibitor (Roche). For analysis of hippocampal tissue, the
FIGURE 1 | Extracting synaptic vesicle (SV) exo- and endocytic rates by population imaging of vGpH. (A) Snapshots of rat hippocampal neurons (DIV 15)
expressing vGpH and stimulated with two consecutive trains of action potentials (APs; 300 APs, 10 Hz) before (first two images) and after (last two images) addition
of Bafilomycin (Baf). Time is in sec. scale bar: 10 µm. (B) Automated segmentation of active boutons based on vGpH responses after the first stimulation.
(C) Individual (gray) and average vGpH (black) traces derived from segmented boutons. Shaded errorbar indicates 95% confidence interval. (D) Measurement of exo-
and endocytic rates based on vGpH responses before and after Baf (see text).
hippocampi from P10 mice were harvested and homogenized from −50 to 40 mV and fitted with a modified Boltzmann
in RIPA buffer using a Dounce tissue homogenizer. Lysates equation:
were cleared by centrifugation and boiled in Laemmli V1/2act −V
sample buffer. Equal amount of total proteins were loaded. I = Gmax (Erev − V)/ 1 + e kact
Samples were then analyzed by SDS-PAGE, transferred onto
nitrocellulose membranes, probed with appropriate primary and
HRP-labeled secondary antibodies and revealed by enhanced where, I = current density (in pA/pF), Gmax = maximum
chemiluminescence. conductance (in nS/pF), Erev = reversal potential, V = measured
potential, V1/2act = midpoint voltage for current activation, and
kact = the slope factor.
Whole-Cell Patch Clamp Recordings of We used a tail protocol to measure current density; cells were
Cav2.1 and Cav2.2 Currents depolarized using 10 mV voltage steps, from −60 to 60 mV.
Cav2.1 or Cav2.2, the auxiliary subunits (GFP-β2a and α2 δ1 ) Following depolarization, tail currents were evoked with a 10 ms
and DISC1 were transiently transfected in HEK 293 cells using pulse at −50 mV. The data were fitted with single Boltzmann
the calcium phosphate method (Huang et al., 2012). Whole-cell equation:
patch-clamp recordings were performed within 36–72 h after V1/2inact −V
transfection. The external solution contained (in mM) 10 HEPES,
I = I min + (Imax − I min )/ 1 + e kinact
140 TEA-MeSO3 and 5 BaCl2 (pH 7.4, 300–310 mOsm). The
glass pipette solution was backfilled with pipette solution (in
mM) 10 HEPES, 5 CsCl, 138 Cs-MeSO3 , 0.5 EGTA, 1 MgCl2 , 2 where, Imax and I min = maximal and minimal current
mg/ml MgATP (pH 7.3, 290–300 mOsm). HEK 293 cells were respectively, V1/2inact = the half-maximal voltage for current
held at −90 mV using the Axopatch 700B amplifier (Axon inactivation, kinact = slope of inactivation curve.
Instruments). The series resistance for all recordings was less
than 5 MΩ; 70–80% compensation on serial resistance and Statistics
cell membrane capacitance were applied. A P/4 protocol was The experimental design of this study implies that data
used to subtract leakage current. All recordings were obtained collected at individual synaptic terminals are clustered according
with an Axon Digidata 1440A data acquisition system, sampled to neuron preparations and imaged fields. Galbraith et al.
at 5–50 kHz and low pass-filtered at 1 kHz or 6 kHz. The (2010) demonstrated that such clustering can adversely affect
I-V curves were obtained from 10 mV voltage-steps ranging statistical inference when not accounted for. In order to
probe for clustering effects (i.e., intra-field correlations), we to a stimulation paradigm that consists of two consecutive
tested whether bouton responses significantly vary from field trains of APs (Fernandez-Alfonso and Ryan, 2004; Burrone
to field. For this, we conducted Wald tests of the null et al., 2006). The amplitude of the first vGpH transient is
hypothesis of no difference in mean outcome across fields governed by the relative rates of SV exo- and endocytosis
within each condition, which revealed (p-value < 0.0001) during stimulation (Figure 1C). To separate the contributions
strong intra-field correlation. To account for these correlations, of exo- and endocytosis, we blocked SV re-acidification with
we performed our statistical analysis in the framework of the vacuolar H+ ATPase inhibitor Bafilomycin (Baf) during
linear mixed models (Laird and Ware, 1982) with normally- the second stimulation, which allows selective measurement
distributed random field effects and preparation fixed effects. of SV exocytosis (Figure 1C). To account for cell-to-cell
In experiments involving one genetically-modified condition variation in vGpH expression, we normalized each trace
and one control group, we tested the null hypothesis of based on the vGpH signal measured after NH4 Cl addition
no difference between the mean outcome of the groups via (Figure 1C). The endocytic component of SV cycling was then
2-sample t-tests. In experiments involving two genetically- computed by subtracting the first vGpH transient from the
modified conditions and one control group, we tested the vGpH response after Baf treatment (Figure 1D). Synaptic
null hypothesis of no difference between the mean outcome traces with baseline drifts between the first and second
of each group and the control group jointly using Wald stimulation or after Baf addition were excluded from the
tests. All tests were performed at the 5% level of statistical analysis (see ‘‘Materials and Methods’’ Section). Exocytosis
significance and carried out using the statistical software Stata largely dominates at the onset of stimulation, while endocytosis
version 13.2. Because our statistical approach (linear mixed kicks in towards the end of the AP train. Exo- and endocytic
model) is not a standard practice when analyzing synaptic rates were finally obtained by linear fitting of the exo- and
properties, we compared the p-values obtained with our method endo traces (Figure 1D). Exocytic rates are about three
with those measured by the more common field averaging times higher than endocytic rates under this stimulation
approach, where the information of a field is collapsed to protocol. To evaluate the variability of bouton responses
a single independent observation by taking the mean of across fields and neuron preparations, we compiled exocytic
bouton responses. Both methods are valid and gave comparable rates from six independent experiments (Supplementary
results (Supplementary Table T1), although the field averaging Figure S1). Statistical analysis of these responses revealed
approach is less statistically efficient as it diminishes the substantial field to field variation (see ‘‘Materials and Methods’’
information that can be obtained from the data by reducing Section), implying that the information collected from each
individual measurements in a field to one observation (Galbraith bouton does not constitute an independent measurement.
et al., 2010). To compare synaptic responses in different fields and
conditions we used a statistical approach that accounts for
correlations within fields and variations in neuron preparations
RESULTS (see ‘‘Materials and Methods’’ Section and Supplementary
Table T1).
DISC1 Loss-of-Function Slows Down SV We then silenced the DISC1 gene using a published shRNA
Cycling sequence (shRNA-E), targeting exon 2 (Duan et al., 2007) and
We opted for an imaging approach based on the synaptic a new shRNA sequence (shRNA-A) targeting exon 9. These
tracer vGpH to explore the impact of DISC1 on the SV cycle. two shRNAs target regions of the DISC1 gene that are fully
vGpH consists of the pH-sensitive GFP variant pHluorin conserved in rodents and are thus expected to silence rat
(Miesenbock et al., 1998) fused to the SV-resident glutamate and mouse DISC1 with the same efficiency. To probe the
transporter vGlut1. pHluorin, which faces the acidic lumen effect of these shRNAs on DISC1 protein levels, we used an
of SVs, undergoes a ∼20-fold increase in fluorescence antibody raised against the C-terminus of mouse DISC1, which
intensity when exposed to the neutral pH of the extracellular has previously been validated in a DISC1 KO mouse (Kuroda
milieu after membrane fusion. (Sankaranarayanan et al., et al., 2011). In our hands, this antibody poorly reacts with rat
2000). Following glutamate discharge and vGpH re-uptake, DISC1 (not shown). In mouse hippocampal neurons, however,
SVs are rapidly re-acidified and vGpH fluorescence is it detects one major protein at ∼100 kDa corresponding to the
quenched. This property has made vGpH a valuable tool predicted full length DISC1 (Kuroda et al., 2011) and which is
to monitor both exo- and endocytosis of SVs at single substantially reduced in both shRNA-E and -A transduced cells
synapses. (Figure 2A).
Due to the inherent variability in the properties of individual Next, we probed the impact of these shRNAs on the
synaptic boutons (Ariel et al., 2013) we measured SV cycling SV cycle in rat hippocampal neurons. We initially chose
in large synapse populations. For this, we developed an image to work with rat, rather than mouse neurons, because we
analysis algorithm that identifies all responding boutons in can harvest substantially more cells from a rat embryonic
a given field (Figures 1A,B), and imaged 16–18 fields from litter. vGpH signals in shRNA-E and -A expressing neurons
at six independent rat hippocampal neuron cultures, yielding display a slower rise and lower amplitude during the first
close to a thousand synaptic boutons for each condition. We and second stimulation, relative to control cells expressing a
employed this algorithm to monitor SV cycling in response scramble shRNA (Figure 2B). Analysis of exocytosis shows
FIGURE 2 | DISC1 silencing by RNAi slows down SV exocytosis. (A) Immunoblot analysis of DISC1 in mouse hippocampal neurons (DIV 8) transduced with
scramble, DISC1-A and DISC1-E shRNAs. (B) Average vGpH traces derived from neurons expressing scr (1204 boutons, 18 fields), -E (960 boutons, 18 fields) and
-A (924 boutons, 16 fields) shRNAs from six independent experiments. (C,D) Analysis of SV exocytosis after Baf treatment. (C) Exocytic profiles of neurons
expressing the indicated shRNAs. (D) Boxplot of exocytic rates. (E,F) Analysis of SV endocytosis. (E) Endocytic profiles. (F) Boxplot of endocytic rates. (G) Exocytic
profile of cells expressing the indicated shRNAs in response to a stimulation (1200 APs, 10 Hz) that depletes the total releasable pool.
substantial synapse-to-synapse variation in all conditions, but display SV cycling defects that are remarkably similar to
reveals a marked decrease in exocytic rates (Figures 2C,D) and these observed with RNAi. The rates and amplitudes of
amplitudes (Figure 2C) in both shRNA-E and -A expressing exocytic responses were substantially reduced relative to wt
neurons. DISC1 silencing also resulted in slightly slower cells (Figures 3B–D), while endocytic rates were marginally,
SV endocytosis, although this effect did not reach statistical but not significantly diminished (Figures 3E,F). Together, these
significance (Figures 2E,F). results show that both genetic ablation and RNAi knockdown
Reduced amplitude of the exocytic response prompted us of DISC1 selectively disrupts SV exocytosis at glutamatergic
to test whether DISC1 also regulates the size of the total synapses with no detectable impact on the total releasable
releasable pool. For this, we applied a stimulus strong enough pool.
(1200 APs at 10 Hz) to maximally deplete releasable SVs
from presynaptic terminals (Ariel and Ryan, 2010). Under this
stimulation regime, vGpH responses reached the same plateau DISC1 Stimulates Ca2+ Influx and
in DISC1-silenced and control neurons, albeit with different Regulates Cav2.2-Dependent SV Release
kinetics, arguing against an effect on the total releasable pool but Because SV exocytosis is initiated by Ca2+ influx through
confirming the stimulatory function of DISC1 in SV exocytosis VGCCs, we measured evoked Ca2+ transients at presynaptic
(Figure 2G). terminals using the SV-targeted Ca2+ sensor SyGC3 (Li
To rule out the possibility of RNAi off-target effects—shRNA- et al., 2011). Ca2+ signals evoked by 300 APs (10 Hz)
E has recently been suggested to inhibit neuron migration or 200 APs (20 Hz) show a rapid rise and partial decay
in the developing cortex independently of DISC1 (Tsuboi during the stimulus (Figures 4A,B,D,E), as observed
et al., 2015)—we examined SV cycling in a DISC1 KO by others (Li et al., 2011). These Ca2+ transients were
mouse that lacks exons 2 and 3 of the DISC1 gene (Kuroda significantly reduced in shRNA-E and -A expressing
et al., 2011). Homozygous DISC1∆2–3/∆2–3 mice show no neurons, under the same stimulation conditions used
detectable levels of the major isoform of DISC1 (Figure 3A). for vGpH measurements (Figures 4A,C), or in response
Hippocampal neurons derived from DISC1∆2–3/∆2–3 mice to a higher frequency stimulus (Figure 4B). Similarly,
FIGURE 3 | Attenuated SV release in hippocampal neurons from DISC1∆2–3/∆2–3 mice. (A) Immunoblot analysis of DISC1 in hippocampal lysates prepared
from P10 DISC1∆2–3/∆2–3 , DISC1wt/∆ 2–3 and DISC1wt/wt mice, confirming the ablation of full-length DISC1 (∼100 kD). (B) Average vGpH traces in DISC1wt/wt
(575 boutons, 7 fields) and DISC1∆2–3/∆2–3 (682 boutons, 9 fields) neurons in response to two consecutive trains of APs and obtained from two independent
experiments. (C,D) Analysis of SV exocytosis from vGpH responses after Baf. Average kinetics (C) and rates (D) of vGpH exocytic responses. (E,F) Analysis of SV
endocytosis. Average kinetics (E) and rates (F) of vGpH endocytic responses.
DISC1∆2–3/∆2–3 neurons show lowered Ca2+ transients channel-specific toxins to determine the contribution of
compared to wt cells (Figures 4D–F). We were concerned each subtype to SV release in hippocampal neurons. ω-
that, under such prolonged stimulations, Ca2+ concentration Conotoxin GVIA (an N-type blocker) reduced SV exocytic
in terminals might reach levels that partially saturate SyGC3. rates by 73 ± 5% (Figures 5A,B), while ω-Agatoxin TK
Ca2+ signals were therefore also examined in response to (a P/Q-type blocker) resulted in a 42 ± 3% decrease in
shorter stimulations (20 APs, 20 Hz) under conditions SV exocytic rates (Figures 5C,D), in a good agreement
well below SyGC3 saturation (Tian et al., 2009; Akerboom with a recent study (Ariel et al., 2013). Note that in these
et al., 2012). A clear decrease in the amplitude of Ca2+ experiments, exocytic rates were approximated by measuring
transients was also observed in DISC1-silenced neurons the initial slope of the vGpH response in the absence of Baf
under this stimulation regime (Figure 4G). Together, these (see Figure 1D). Importantly, blockade of Cav2.2 almost
Ca2+ imaging data hint at a Ca2+ influx defect in DISC1- completely occluded the effect of DISC1 silencing on SV
inactivated cells, although abnormal clearance of Ca2+ from the exocytosis (Figures 5A,B,E). Blockade of Cav2.1, on the
presynaptic terminal could also be involved (see ‘‘Discussion’’ other hand, slightly increased the inhibitory impact of
Section). DISC1 knockdown on SV release rates (Figures 5C–E).
To determine whether reduced Ca2+ influx is the primary We conclude from these results that DISC1 selectively
cause of SV cycling defects, we attempted to rescue SV exocytosis stimulates Cav2.2-dependent SV release at hippocampal
in DISC1-silenced neurons by elevating extracellular Ca2+ synapses.
concentration. Shifting extracellular Ca2+ concentration from 2 We then asked whether DISC1 regulates Cav2.2 activity
to 4 mM increases AP-induced Ca2+ entry (Ariel and Ryan, 2010) by controlling its presynaptic localization and/or abundance.
and restored the vGpH response (Figure 4H), in line with a role Co-localization studies of endogenous Cav2.2 with the active
of DISC1 in facilitating Ca2+ influx. zone marker bassoon revealed extensive presence of Cav2.2 in
At central synapses, the P/Q-type (Cav2.1) and N-type synaptic boutons, both in control and DISC1-silenced neurons
(Cav2.2) Ca2+ channels are the main sources of Ca2+ for (Supplementary Figures S2A–C). We found no significant
initiation of SV exocytosis (Catterall et al., 2013). We used difference between these two groups in the fraction of boutons
FIGURE 4 | DISC1 loss-of-function reduces evoked Ca2+ transients at nerve terminals. SyGC3 imaging in rat hippocampal neurons (DIV14–16) in response
two different trains of APs. (A) Average Ca2+ transients in neurons expressing scr (1052 boutons, 15 fields, 5 experiments (exps)), DISC1-E (1836 boutons, 9 fields,
3 exps) and -A (754 boutons, 11 fields, 3 exps) shRNAs, in response to 300 APs, 10 Hz. The DISC1-E and DISC-A groups are significantly different than the scr
group (p = 0.0057). (B) Average Ca2+ transients in neurons expressing scr (861 boutons, 9 fields, 5 exps), DISC1-E (2118 boutons, 7 fields, 3 exps) and -A (1549
boutons, 9 fields, 3 exps) shRNAs, in response to 200 APs, 20 Hz. (C) Cumulative probability of SyGC3 peak intensity from individual boutons corresponding to
(A). (D) Average Ca2+ transients in DISC1 wt/wt (1090 boutons, 11 fields, 3 exps) and DISC1∆2–3/∆2–3 (727 boutons, 9 fields, 3 exps) neurons, in response to 300
APs, 10 Hz. The DISC1 wt/wt and DISC1∆2–3/∆2–3 groups are statistically different (p = 0.0182). (E) Average Ca2+ transients in DISC1 wt/wt (1365 boutons, 9 fields, 3
exps) and DISC1∆2–3/∆2–3 (1233 boutons, 8 fields, 3 exps) neurons, in response to 200 APs, 20 Hz. (F) Cumulative probability of SyGC3 peak intensity from
individual boutons corresponding to (D). (G) Average Ca2+ transients in neurons expressing scr (290 boutons, 6 fields, 2 exps), and DISC1-E (390 boutons, 6 fields,
2 exps) shRNAs, in response to 20 APs, 20 Hz. The scr and DISC1-E groups are statistically different (p = 0.0036). (H) Average vGpH traces in scr (180 boutons, 5
fields, 2 exps) and DISC1-E (488 boutons, 6 fields, 2 exps) shRNA-expressing neurons during two consecutive trains of APs (300 AP, 10 Hz) in the presence of 2 or
4 mM extracellular Ca2+ .
containing Cav2.2, or in the intensity of Cav2.2 staining in We finally examined the effect of DISC1 on Cav2 currents
presynaptic terminals (Supplementary Figures S2B,C). Nor by whole cell patch-clamp recordings. Co-transfection of
did we find evidence for altered presynaptic localization and recombinant DISC1 (Figure 6A) in HEK293T cells together
intensity of Cav2.1 (Supplementary Figures S2D–F or reduced with Cav2.2 and its auxiliary subunits (see ‘‘Materials and
density of presynaptic terminals (Supplementary Figure S2G) Methods’’ Section) resulted in a 27% increase in Cav2.2
in DISC1 knockdown neurons. Thus, synaptic targeting and peak current density (Figures 6B,C). Likewise, tail current
abundance of Cav2 channels does not seem to be regulated by density of Cav2.2 was substantially elevated in DISC1-
DISC1. overexpressing cells (Figures 6D,E). Current density can be
FIGURE 5 | DISC1 regulates Cav2.2-dependent SV exocytosis. (A) Average vGpH traces in scr (101 boutons, 6 fields, 2 exps) and DISC1-E (87 boutons,
5 fields, 2 exps) shRNA-expressing neurons during consecutive trains of APs (300 AP, 10 Hz) in the absence or presence of the Cav2.2 blocker ω-Conotoxin GVIA
(125 nM). (B) Boxplot of SV exocytic rates before and after ω-Conotoxin GVIA application. Exocytic rates were measured by linear fitting of the first six time points of
the vGpH response. (C) Average vGpH traces in scr (1294 boutons, 9 fields, 3 exps) and DISC1-E (1282 boutons, 9 fields, 3 exps) shRNA-expressing neurons during
consecutive trains of APs (300 AP, 10 Hz) in the absence or presence of the Cav2.1 blocker ω-Agatoxin TK (125 nM). (D) Boxplot of SV exocytic rates before and
after ω-Agatoxin TK application. (E) Table showing the percentage of inhibition of exocytosis rate by DISC1 knockdown before and after Cav2.2- or Cav2.1 blockade.
affected by a change in channel gating or by the number that restricts the activity of DISC1 to Cav2.2-dependent SV
of channels at the cell surface. To test the first possibility, exocytosis.
we examined the voltage-dependent properties of Cav2.2
current activation and saw no difference between control DISCUSSION
and DISC1-overexpressing cells (Figure 6F). This suggests
that DISC1 promotes surface delivery and/or stabilize We used here two independent gene-targeting approaches
surface expression of Cav2.2. Recording of Cav2.1 currents together with large-scale imaging of presynaptic function to
revealed a similar, voltage-independent, potentiation effect determine the impact of the mental disease gene DISC1
of DISC1 (Figures 6G–K). Thus, DISC1 equally augments on the glutamate release machinery. Our results show that
both Cav2.2 and Cav2.1 currents in this heterologous DISC1 accelerates SV exocytosis and thus enhances presynaptic
system, presumably by increasing surface expression of performance. This boosting effect is mediated by N-type Ca2+
these Ca2+ channels. These results also imply the presence channels, establishing the first link between DISC1 and VGCC
of an additional layer of regulation in hippocampal neurons activity.
FIGURE 6 | DISC1 enhances Cav2.2 and Cav2.1 currents. (A) Western blot showing expression of ectopic (human) DISC1 in HEK293 cells. (B,C) Cav2.2
Current-voltage (I-V) curves for hDISC1-expressing and control cells. (B) Stimulation protocol and individual current responses shown at three different voltages
(−30, 0 and 30 mV) (C). Average I-V plots for hDISC1 (peak = 54.2 ± 4.5 pA/pF, 13 cells) and control (peak = 39.2 ± 4.9 pA/pF, 12 cells), p = 0.033. (D–F) Cav2.2
activation curves in response to the tail protocol. (D) Illustration of the tail protocol and individual tail currents measured at −50 mV after three different voltage steps
(−20, 0 and 40 mV). (E) Average current density based on tail currents for DISC1 (164.3 ± 7.3 pA/pF, 12 cells) and control (110.5 ± 6.3 pA/pF, 12 cells), ∗ p < 0.001.
(F) Normalized activation curve from tail currents showing no significant difference between DISC1 (V50 : −6.98 ± 3.43 mV, 12 cells) and control (V50 : 0.86 ± 3.62
mV, 12 cells), p = 0.13. (G,H) Cav2.1 Current-voltage (I-V) curves for hDISC1-expressing and control cells. (G) Stimulation protocol and individual current responses
shown at three different voltages (−30, 0 and 30 mV). (H) Average I-V plots for hDISC1 (peak = 79.9 ± 8.3 pA/pF, 15 cells) and control (peak = 56.3 ± 7.4 pA/pF,
13 cells), p = 0.046. (I–K) Cav2.1 activation curves in response to the tail protocol. (I) Illustration of the tail protocol and individual tail currents measured at −50 mV
after three different voltage steps (−20, 0 and 40 mV). (J) Average current density based on tail currents for DISC1 (156.3 ± 5.7 pA/pF, 25 cells) and control
(123.2 ± 7.4 pA/pF, 17 cells), ∗ p = 0.003. (K) Normalized activation curve from tail currents showing no significant difference between DISC1 (V50 : −5.39 ± 0.49
mV, 24 cells) and control (V50 : −4.41 ± 0.93 mV, 17 cells), p = 0.31.
DISC1 inactivation in hippocampal neurons results in The efficacy of neurotransmitter release determines not only
decreased Ca2+ transients at hippocampal terminals in response the strength of synaptic excitation, but also dictates various forms
to AP firing, consistent with a deficit in Ca2+ entry. Because of short-term plasticity (Abbott and Regehr, 2004), suggesting
these Ca2+ signals are shaped by both Ca2+ entry and broad functions of DISC1 in synaptic computation and neural
clearance processes, we cannot exclude a role of DISC1 in circuit performance. Although we have not directly measured
regulating Ca2+ extrusion or buffering. A stimulatory effect release probability (Pr ), the impact of DISC1 on Ca2+ entry
of DISC1 on Ca2+ entry is further supported, however, by and SV cycling suggests a positive influence on Pr (see Maher
its positive impact on Cav2.2 and Cav2.1 currents, the two and Loturco, 2012). Because synaptic facilitation and depression
main VGCCs underlying the initiation of neurotransmission depend largely on the initial Pr (high Pr favors depression while
at hippocampal synapses. How DISC1 enhances Cav2 currents low Pr favors facilitation), our results predict altered short-
is unclear. The lack of an effect on gating suggests that term plasticity in DISC1-deficient neural circuits. Of interest,
DISC1 promotes surface expression of these Ca2+ channels, abnormal short-term plasticity has been associated with deficits
although we cannot exclude a role of DISC1 in regulating in moment-to-moment information processing and working
single-channel conductance or open probability. Our immuno- memory, two hallmarks of schizophrenia (Crabtree and Gogos,
localization studies revealed no detectable impact of DISC1 2014). Reduced efficacy in glutamate release may also be relevant
on presynaptic abundance of Cav2 channels. This does not for forms of long-term potentiation (LTP) that have clear
rule out, however, the possibility that DISC1 regulates the presynaptic components, such as LTP at the perforant path-
number of functional Ca2+ channels at the surface of the granule cell synapse in the dentate gyrus (Errington et al.,
terminal. 1987). Notably, this form of LTP is impaired in DISC1∆2–3/∆2–3
Although overexpression of DISC1 enhances both Cav2.1 mice—a stronger tetanic stimulus is required for the expression
and Cav2.2 currents in a heterologous system, our findings of LTP (Kuroda et al., 2011). Our findings suggest that this
point to a selective influence of DISC1 on Cav2.2-dependent LTP deficit could be due to a failure of the presynaptic
SV exocytosis. How is this selectivity achieved? It is possible terminal to undergo activity-dependent increase in release
that both Cav2.1 and Cav2.2 compete for DISC1 regulation at probability.
presynaptic terminals. The dominant contribution of Cav2.2 to Several genes encoding VGCC subunits, including
SV release probably results (at least in part) from increased CACNA1C, CACNB2 and CACNA1I have repeatedly been
levels of Cav2.2 (relative to Cav2.1) at hippocampal terminals associated with schizophrenia and other psychiatric disorders
(Ariel et al., 2013). Thus, preferential regulation of Cav2.2 by (Ferreira et al., 2008; Cross-Disorder Group of the Psychiatric
endogenous levels of DISC1 combined with greater abundance Genomics, 2013; Hamshere et al., 2013; Ripke et al., 2013;
of Cav2.2 may explain why DISC1-dependent SV exocytosis Schizophrenia Working Group of the Psychiatric Genomics,
appears to be exclusively mediated by Cav2.2. Specificity could 2014). Although CACNA1A (Cav2.1) and CACNA1B (Cav2.2)
also arise, however, from functions of DISC1 downstream are not typically associated with risk loci, RIM1 (also called
of Ca2+ entry. For example, DISC1 could be involved in RIMS1)—a presynaptic scaffold that regulates density of
positioning Cav2.2 in close proximity to the SV release P/Q- and N-type Ca2+ channels and SV docking at the
machinery at the active zone, a process that dictates speed and active zone (Han et al., 2011)—was recently identified as a
fidelity of neurotransmission. While the role of Ca2+ entry candidate gene for schizophrenia in the largest GWAS study
in SV exocytosis is undisputed, its impact on SV endocytosis conducted to date (Schizophrenia Working Group of the
remains somewhat controversial. Recent evidence suggests, Psychiatric Genomics, 2014). Together these findings point
however, that Ca2+ couples rates of SV exo- and endocytosis to neurotransmitter release as a central process targeted in
and optimizes endocytic rates during AP bursts (Armbruster schizophrenia.
et al., 2013). We observed reduced endocytic rates in both In conclusion, our results shed light on a novel
DISC1-silenced and DISC1∆2–3/∆2–3 neurons. Although this mechanism by which a major susceptibility gene for mental
effect was small and did not reach statistical significance, illness enhances the efficacy of glutamate release, and
our results suggest that the influence of DISC1 on Cav2.2 provide further support for a central role of glutamate
impacts both exo- and endocytosis of SVs at hippocampal neurotransmission in schizophrenia and other major mood
terminals. disorders.
Our findings extend on two recent studies implicating
DISC1 in presynaptic function. In glutamatergic neurons AUTHOR CONTRIBUTIONS
differentiated from human iPS cells and derived from members
of a family with a frameshift mutation in DISC1, deficits WT and JVT performed and analyzed all imaging and
in SV release were observed after KCl-induced depolarization biochemical experiments. QL performed and analyzed whole-cell
(Wen et al., 2014). In a separate report, RNAi knockdown patch-clamp recordings. MB carried out the statistical analysis.
of DISC1 in layer 2/3 neocortical neurons increases paired- KBL made the original observation implicating DISC1 in SV
pulse facilitation and appears to reduce probability of glutamate cycling. KKu and KKa made the DISC1 ∆2–3 mouse and
release (Maher and Loturco, 2012). Collectively, these results the DISC1 C-terminus antibody. WT and MF wrote Matlab
clearly identify DISC1 as a positive modulator of glutamate scripts. TWS and MF supervised the project. MF wrote the
release. article.
REFERENCES Duan, X., Chang, J. H., Ge, S., Faulkner, R. L., Kim, J. Y., Kitabatake, Y., et al.
(2007). Disrupted-In-Schizophrenia 1 regulates integration of newly generated
Abbott, L. F., and Regehr, W. G. (2004). Synaptic computation. Nature 431, neurons in the adult brain. Cell 130, 1146–1158. doi: 10.1016/j.cell.2007.
796–803. doi: 10.1038/nature03010 07.010
Akerboom, J., Chen, T. W., Wardill, T. J., Tian, L., Marvin, J. S., Mutlu, Ekelund, J., Hennah, W., Hiekkalinna, T., Parker, A., Meyer, J., Lonnqvist, J., et al.
S., et al. (2012). Optimization of a GCaMP calcium indicator for neural (2004). Replication of 1q42 linkage in Finnish schizophrenia pedigrees. Mol.
activity imaging. J. Neurosci. 32, 13819–13840. doi: 10.1523/jneurosci.2601- Psychiatry 9, 1037–1041. doi: 10.1038/sj.mp.4001536
12.2012 Ekelund, J., Hovatta, I., Parker, A., Paunio, T., Varilo, T., Martin, R., et al. (2001).
Ariel, P., Hoppa, M. B., and Ryan, T. A. (2013). Intrinsic variability in Pv, Chromosome 1 loci in Finnish schizophrenia families. Hum. Mol. Genet. 10,
RRP size, Ca2+ channel repertoire and presynaptic potentiation in individual 1611–1617. doi: 10.1093/hmg/10.15.1611
synaptic boutons. Front. Synaptic Neurosci. 4:9. doi: 10.3389/fnsyn.2012. Errington, M. L., Lynch, M. A., and Bliss, T. V. (1987). Long-term potentiation in
00009 the dentate gyrus: induction and increased glutamate release are blocked by
Ariel, P., and Ryan, T. A. (2010). Optical mapping of release properties in synapses. D(-)aminophosphonovalerate. Neuroscience 20, 279–284. doi: 10.1016/0306-
Front. Neural Circuits 4:18. doi: 10.3389/fncir.2010.00018 4522(87)90019-4
Armbruster, M., Messa, M., Ferguson, S. M., De Camilli, P., and Ryan, T. A. Fernandez-Alfonso, T., and Ryan, T. A. (2004). The kinetics of synaptic vesicle
(2013). Dynamin phosphorylation controls optimization of endocytosis pool depletion at CNS synaptic terminals. Neuron 41, 943–953. doi: 10.
for brief action potential bursts. Elife 2:e00845. doi: 10.7554/elife. 1016/s0896-6273(04)00113-8
00845 Ferreira, M. A., O’donovan, M. C., Meng, Y. A., Jones, I. R., Ruderfer, D. M.,
Burrone, J., Li, Z., and Murthy, V. N. (2006). Studying vesicle cycling Jones, L., et al. (2008). Collaborative genome-wide association analysis supports
in presynaptic terminals using the genetically encoded probe a role for ANK3 and CACNA1C in bipolar disorder. Nat. Genet. 40, 1056–1058.
synaptopHluorin. Nat. Protoc. 1, 2970–2978. doi: 10.1038/nprot. doi: 10.1038/ng.209
2006.449 Fromer, M., Pocklington, A. J., Kavanagh, D. H., Williams, H. J., Dwyer,
Catterall, W. A., Leal, K., and Nanou, E. (2013). Calcium channels and short- S., Gormley, P., et al. (2014). De novo mutations in schizophrenia
term synaptic plasticity. J. Biol. Chem. 288, 10742–10749. doi: 10.1074/jbc.r112. implicate synaptic networks. Nature 506, 179–184. doi: 10.1038/nature
411645 12929
Chubb, J. E., Bradshaw, N. J., Soares, D. C., Porteous, D. J., and Millar, J. K. (2008). Galbraith, S., Daniel, J. A., and Vissel, B. (2010). A study of clustered data and
The DISC locus in psychiatric illness. Mol. Psychiatry 13, 36–64. doi: 10.1038/sj. approaches to its analysis. J. Neurosci. 30, 10601–10608. doi: 10.1523/jneurosci.
mp.4002106 0362-10.2010
Clapcote, S. J., Lipina, T. V., Millar, J. K., Mackie, S., Christie, S., Ogawa, F., et al. Garcia-Alvarez, G., Lu, B., Yap, K. A., Wong, L. C., Thevathasan, J. V., Lim,
(2007). Behavioral phenotypes of Disc1 missense mutations in mice. Neuron L., et al. (2015). STIM2 regulates PKA-dependent phosphorylation and
54, 387–402. doi: 10.1016/j.neuron.2007.04.015 trafficking of AMPARs. Mol. Biol. Cell 26, 1141–1159. doi: 10.1091/mbc.e14-
Crabtree, G. W., and Gogos, J. A. (2014). Synaptic plasticity, neural circuits 07-1222
and the emerging role of altered short-term information processing in Glessner, J. T., Reilly, M. P., Kim, C. E., Takahashi, N., Albano, A., Hou, C.,
schizophrenia. Front. Synaptic Neurosci. 6:28. doi: 10.3389/fnsyn.2014. et al. (2010). Strong synaptic transmission impact by copy number variations
00028 in schizophrenia. Proc. Natl. Acad. Sci. U S A 107, 10584–10589. doi: 10.
Cross-Disorder Group of the Psychiatric Genomics, C. (2013). Identification 1073/pnas.1000274107
of risk loci with shared effects on five major psychiatric disorders: a Hamshere, M. L., Walters, J. T., Smith, R., Richards, A. L., Green, E., Grozeva,
genome-wide analysis. Lancet 381, 1371–1379. doi: 10.1016/s0140-6736(12) D., et al. (2013). Genome-wide significant associations in schizophrenia to
62129-1 ITIH3/4, CACNA1C and SDCCAG8 and extensive replication of associations
Cross-Disorder Group of the Psychiatric Genomics, C., Lee, S. H., Ripke, S., Neale, reported by the Schizophrenia PGC. Mol. Psychiatry 18, 708–712. doi: 10.
B. M., Faraone, S. V., Purcell, S. M., et al. (2013). Genetic relationship between 1038/mp.2012.67
five psychiatric disorders estimated from genome-wide SNPs. Nat. Genet. 45, Han, Y., Kaeser, P. S., Sudhof, T. C., and Schneggenburger, R. (2011).
984–994. doi: 10.1038/ng.2711 RIM determines Ca2+ channel density and vesicle docking at the
presynaptic active zone. Neuron 69, 304–316. doi: 10.1016/j.neuron. Millar, J. K., Wilson-Annan, J. C., Anderson, S., Christie, S., Taylor, M. S.,
2010.12.014 Semple, C. A., et al. (2000). Disruption of two novel genes by a translocation
Hikida, T., Jaaro-Peled, H., Seshadri, S., Oishi, K., Hookway, C., Kong, S., et al. co-segregating with schizophrenia. Hum. Mol. Genet. 9, 1415–1423. doi: 10.
(2007). Dominant-negative DISC1 transgenic mice display schizophrenia- 1093/hmg/9.9.1415
associated phenotypes detected by measures translatable to humans. Proc. Moskvina, V., Craddock, N., Holmans, P., Nikolov, I., Pahwa, J. S., Green, E., et al.
Natl. Acad. Sci. U S A 104, 14501–14506. doi: 10.1073/pnas.07047 (2009). Gene-wide analyses of genome-wide association data sets: evidence
74104 for multiple common risk alleles for schizophrenia and bipolar disorder and
Howes, O., Mccutcheon, R., and Stone, J. (2015). Glutamate and dopamine in for overlap in genetic risk. Mol. Psychiatry 14, 252–260. doi: 10.1038/mp.
schizophrenia: an update for the 21st century. J. Psychopharmacol. 29, 97–115. 2008.133
doi: 10.1177/0269881114563634 Niwa, M., Kamiya, A., Murai, R., Kubo, K., Gruber, A. J., Tomita, K.,
Huang, H., Tan, B. Z., Shen, Y., Tao, J., Jiang, F., Sung, Y. Y., et al. (2012). et al. (2010). Knockdown of DISC1 by in utero gene transfer disturbs
RNA editing of the IQ domain in Ca(v)1.3 channels modulates their Ca2+ - postnatal dopaminergic maturation in the frontal cortex and leads to
dependent inactivation. Neuron 73, 304–316. doi: 10.1016/j.neuron.2011. adult behavioral deficits. Neuron 65, 480–489. doi: 10.1016/j.neuron.2010.
11.022 01.019
Huettner, J. E., and Baughman, R. W. (1986). Primary culture of identified Pletnikov, M. V., Ayhan, Y., Nikolskaia, O., Xu, Y., Ovanesov, M. V.,
neurons from the visual cortex of postnatal rats. J. Neurosci. 6, Huang, H., et al. (2008). Inducible expression of mutant human DISC1
3044–3060. in mice is associated with brain and behavioral abnormalities reminiscent
International Schizophrenia, C., Purcell, S. M., Wray, N. R., Stone, J. L., Visscher, of schizophrenia. Mol. Psychiatry 13, 173–186. doi: 10.1038/sj.mp.40
P. M., O’donovan, M. C., et al. (2009). Common polygenic variation contributes 02079
to risk of schizophrenia and bipolar disorder. Nature 460, 748–752. doi: 10. Poon, V. Y., Goh, C., Voorhoeve, P. M., and Fivaz, M. (2014). High-content
1038/nature08185 imaging of presynaptic assembly. Front. Cell Neurosci. 8:66. doi: 10.3389/fncel.
Kaech, S., and Banker, G. (2006). Culturing hippocampal neurons. Nat. Protoc. 1, 2014.00066
2406–2415. doi: 10.1038/nprot.2006.356 Ripke, S., O’dushlaine, C., Chambert, K., Moran, J. L., Kahler, A. K.,
Kamiya, A., Kubo, K., Tomoda, T., Takaki, M., Youn, R., Ozeki, Y., Akterin, S., et al. (2013). Genome-wide association analysis identifies 13
et al. (2005). A schizophrenia-associated mutation of DISC1 perturbs new risk loci for schizophrenia. Nat. Genet. 45, 1150–1159. doi: 10.1038/
cerebral cortex development. Nat. Cell Biol. 7, 1167–1178. doi: 10.1038/ ng.2742
ncb1328 Sachs, N. A., Sawa, A., Holmes, S. E., Ross, C. A., Delisi, L. E., and Margolis, R. L.
Kilpinen, H., Ylisaukko-Oja, T., Hennah, W., Palo, O. M., Varilo, T., Vanhala, R., (2005). A frameshift mutation in Disrupted in Schizophrenia 1 in an American
et al. (2008). Association of DISC1 with autism and Asperger syndrome. Mol. family with schizophrenia and schizoaffective disorder. Mol. Psychiatry 10,
Psychiatry 13, 187–196. doi: 10.1038/sj.mp.4002031 758–764. doi: 10.1038/sj.mp.4001667
Kubo, K., Tomita, K., Uto, A., Kuroda, K., Seshadri, S., Cohen, J., et al. Sankaranarayanan, S., De Angelis, D., Rothman, J. E., and Ryan, T. A.
(2010). Migration defects by DISC1 knockdown in C57BL/6, 129X1/SvJ (2000). The use of pHluorins for optical measurements of presynaptic
and ICR strains via in utero gene transfer and virus-mediated RNAi. activity. Biophys. J. 79, 2199–2208. doi: 10.1016/s0006-3495(00)
Biochem. Biophys. Res. Commun. 400, 631–637. doi: 10.1016/j.bbrc.2010. 76468-x
08.117 Schizophrenia Working Group of the Psychiatric Genomics, C. (2014). Biological
Kuroda, K., Yamada, S., Tanaka, M., Iizuka, M., Yano, H., Mori, D., et al. (2011). insights from 108 schizophrenia-associated genetic loci. Nature 511, 421–427.
Behavioral alterations associated with targeted disruption of exons 2 and 3 doi: 10.1038/nature13595
of the Disc1 gene in the mouse. Hum. Mol. Genet. 20, 4666–4683. doi: 10. Shen, S., Lang, B., Nakamoto, C., Zhang, F., Pu, J., Kuan, S. L., et al.
1093/hmg/ddr400 (2008). Schizophrenia-related neural and behavioral phenotypes in transgenic
Laird, N. M., and Ware, J. H. (1982). Random-effects models for longitudinal data. mice expressing truncated Disc1. J. Neurosci. 28, 10893–10904. doi: 10.
Biometrics 38, 963–974. 1523/jneurosci.3299-08.2008
Lee, F. H., Fadel, M. P., Preston-Maher, K., Cordes, S. P., Clapcote, S. J., Price, Shinoda, T., Taya, S., Tsuboi, D., Hikita, T., Matsuzawa, R., Kuroda, S.,
D. J., et al. (2011). Disc1 point mutations in mice affect development of et al. (2007). DISC1 regulates neurotrophin-induced axon elongation via
the cerebral cortex. J. Neurosci. 31, 3197–3206. doi: 10.1523/jneurosci.4219- interaction with Grb2. J. Neurosci. 27, 4–14. doi: 10.1523/jneurosci.3825-
10.2011 06.2007
Li, H., Foss, S. M., Dobryy, Y. L., Park, C. K., Hires, S. A., Shaner, N. C., et al. (2011). Singh, K. K., Ge, X., Mao, Y., Drane, L., Meletis, K., Samuels, B. A.,
Concurrent imaging of synaptic vesicle recycling and calcium dynamics. Front. et al. (2010). Dixdc1 is a critical regulator of DISC1 and embryonic
Mol. Neurosci. 4:34. doi: 10.3389/fnmol.2011.00034 cortical development. Neuron 67, 33–48. doi: 10.1016/j.neuron.2010.
Maher, B. J., and Loturco, J. J. (2012). Disrupted-in-schizophrenia (DISC1) 06.002
functions presynaptically at glutamatergic synapses. PLoS One 7:e34053. St Clair, D., Blackwood, D., Muir, W., Carothers, A., Walker, M., Spowart, G.,
doi: 10.1371/journal.pone.0034053 et al. (1990). Association within a family of a balanced autosomal translocation
Mao, Y., Ge, X., Frank, C. L., Madison, J. M., Koehler, A. N., Doud, M. K., with major mental illness. Lancet 336, 13–16. doi: 10.1016/0140-6736(90)
et al. (2009). Disrupted in schizophrenia 1 regulates neuronal progenitor 91520-k
proliferation via modulation of GSK3beta/beta-catenin signaling. Cell 136, Steinecke, A., Gampe, C., Valkova, C., Kaether, C., and Bolz, J. (2012).
1017–1031. doi: 10.1016/j.cell.2008.12.044 Disrupted-in-Schizophrenia 1 (DISC1) is necessary for the correct migration
Miesenbock, G., De Angelis, D. A., and Rothman, J. E. (1998). of cortical interneurons. J. Neurosci. 32, 738–745. doi: 10.1523/jneurosci.5036-
Visualizing secretion and synaptic transmission with pH-sensitive 11.2012
green fluorescent proteins. Nature 394, 192–195. doi: 10.1038/ Thevathasan, J. V., Tan, E., Zheng, H., Lin, Y. C., Li, Y., Inoue, T., et al. (2013). The
28190 small GTPase HRas shapes local PI3K signals through positive feedback and
Millar, J. K., Christie, S., Anderson, S., Lawson, D., Hsiao-Wei Loh, D., Devon, regulates persistent membrane extension in migrating fibroblasts. Mol. Biol.
R. S., et al. (2001). Genomic structure and localisation within a linkage Cell 24, 2228–2237. doi: 10.1091/mbc.E12-12-0905
hotspot of Disrupted In Schizophrenia 1, a gene disrupted by a translocation Tian, L., Hires, S. A., Mao, T., Huber, D., Chiappe, M. E., Chalasani, S. H.,
segregating with schizophrenia. Mol. Psychiatry 6, 173–178. doi: 10.1038/sj.mp. et al. (2009). Imaging neural activity in worms, flies and mice with
4000784 improved GCaMP calcium indicators. Nat. Methods 6, 875–881. doi: 10.1038/
Millar, J. K., Pickard, B. S., Mackie, S., James, R., Christie, S., Buchanan, S. R., nmeth.1398
et al. (2005). DISC1 and PDE4B are interacting genetic factors in schizophrenia Tiscornia, G., Singer, O., and Verma, I. M. (2006). Production and
that regulate cAMP signaling. Science 310, 1187–1191. doi: 10.1126/science. purification of lentiviral vectors. Nat. Protoc. 1, 241–245. doi: 10.1038/nprot.
1112915 2006.37
Tsuboi, D., Kuroda, K., Tanaka, M., Namba, T., Iizuka, Y., Taya, S., Conflict of Interest Statement: The authors declare that the research was
et al. (2015). Disrupted-in-schizophrenia 1 regulates transport of ITPR1 conducted in the absence of any commercial or financial relationships that could
mRNA for synaptic plasticity. Nat. Neurosci. 18, 698–707. doi: 10.1038/ be construed as a potential conflict of interest.
nn.3984
Voglmaier, S. M., Kam, K., Yang, H., Fortin, D. L., Hua, Z., Nicoll, R. A., et al. Copyright © 2016 Tang, Thevathasan, Lin, Lim, Kuroda, Kaibuchi, Bilger, Soong
(2006). Distinct endocytic pathways control the rate and extent of synaptic and Fivaz. This is an open-access article distributed under the terms of the Creative
vesicle protein recycling. Neuron 51, 71–84. doi: 10.1016/j.neuron.2006. Commons Attribution License (CC BY). The use, distribution and reproduction in
05.027 other forums is permitted, provided the original author(s) or licensor are credited
Wen, Z., Nguyen, H. N., Guo, Z., Lalli, M. A., Wang, X., Su, Y., et al. (2014). and that the original publication in this journal is cited, in accordance with accepted
Synaptic dysregulation in a human iPS cell model of mental disorders. Nature academic practice. No use, distribution or reproduction is permitted which does not
515, 414–418. doi: 10.1038/nature13716 comply with these terms.
The largest family of single color genetically encoded calcium overall release probability of the synapse. Whilst this elegant
indicators (GECIs) is the GCaMP family, based on a circularly study was highly informative, the techniques used are limited
permuted GFP linked to calmodulin (CaM) and it’s binding to relatively small numbers of cells and cannot be transferred
peptide M13 (Nakai et al., 2001). The original sensor has to an in vivo situation. With genetically encoded reporters,
undergone multiple rounds of both random and structure-guided visualization of both calcium influx and neurotransmitter release
mutagenesis to improve stability, brightness, signal to noise ratio has been achieved by co-expression of a red-shifted reporter of
and kinetics (Tian et al., 2009; Akerboom et al., 2012; Chen vesicle exocytosis, either VGLUT-mOr2 or sypHTomato, with a
et al., 2013) and several variants have been effectively targeted to green calcium indicator, such as members of the GCaMP family
the neuronal presynaptic terminal by fusion with synaptophysin (Li et al., 2011; Li and Tsien, 2012). However, in this situation
(Dreosti et al., 2009; Li et al., 2011; Nikolaou et al., 2012). To the two probes were expressed separately, resulting in possible
enable multicolor imaging in single cells, red-shifted GECIs have differences in the level of expression of each reporter and in
been developed, including RCaMP (Akerboom et al., 2013) and their subcellular location. Crucially, it also precludes the use of a
R-GECO1 (Zhao Y. et al., 2011), for which a presynaptically ratiometric approach to measuring sypHy responses (Rose et al.,
targeted version (SyRGECO) has also been characterized (Walker 2013), complicating the direct comparison of responses across
et al., 2013). Much like the GCaMPs, red GECIs are constantly cells.
evolving, with new variants recently produced in both RCaMP We have generated a new genetically encoded probe that
and R-GECO families (Inoue et al., 2015; Dana et al., 2016). brings together the measure of calcium and of exocytosis, in
The most widely used reporters of synaptic vesicle exocytosis addition to a ratiometric measure of sypHy signals. To do this
and endocytosis consist of a synaptic vesicle protein fused to we combined sypHy, a well characterized pHluorin sensor of
pHluorin, a pH-sensitive form of GFP (Miesenbock et al., 1998). synaptic vesicle exocytosis (Granseth et al., 2006), with the red-
Crucially in these constructs, pHluorin is localized to the acidic shifted calcium indicator R-GECO1 (Zhao Y. et al., 2011) in a
lumen of synaptic vesicles, so that its fluorescence is quenched single molecule (sypHy-RGECO) to enable concurrent imaging
at rest, but increases by up to 20-fold upon fusion with the of calcium influx and neurotransmitter release. The fixed 1:1 ratio
plasma membrane and exposure to the more basic pH of the of the two reporter molecules allows not only the normalization
extracellular medium (Sankaranarayanan et al., 2000). Several of sypHy signals to baseline or maximum RGECO fluorescence,
variants exist including fusions to the intraluminal domains of but also the use of the ratio of the two responses as a measure
VAMP2, known as synaptopHluorin, (Miesenbock et al., 1998), of the calcium dependence of release at individual synapses.
VGLUT1 (Voglmaier et al., 2006), and synaptophysin, known We systematically varied the number and frequency of action
as sypHy (Granseth et al., 2006). Recently, efforts have been potentials, as well as the external calcium concentrations, and
made to improve the normalization of presynaptic pHluorin measured sypHy and RGECO signals in individual synapses. We
responses, which are complicated by the presence of a fraction of found that while the amplitude of both responses changed, the
the protein at the cell surface, particularly for synaptopHluorin distribution of the ratios remained constant. Interestingly, the
and sypHy. At rest, surface stranded protein dominates the calcium dependence of release was found to vary considerably
baseline fluorescence due to its increased brightness compared between synapses, even within the same cell, suggesting the
to quenched protein localized to synaptic vesicles and is sensitivity of neurotransmitter release to calcium is synapse-
highly variable between boutons, preventing the use of baseline specific. We also combined live sypHy-RGECO imaging with
fluorescence as a normalizing factor for exocytic responses. By immunofluorescence for structural markers and observed that
fusing tdimer2 to the C-terminus of sypHy Rose et al. (2013) the amplitudes of calcium influx and exocytosis at individual
generated ratio-sypHy, in which the red fluorescence provides an synapses were positively correlated with the levels of the active
invariant signal proportional to the expression of the probe rather zone protein RIM but were less well correlated with levels of the
than the surface fraction. The ratio of green to red fluorescence synaptic vesicle calcium sensor synaptotagmin-1.
is fixed and allows normalization of sypHy responses to the
red baseline, simplifying comparisons across cells with different
expression levels and at different depths within a tissue. Finally, MATERIALS AND METHODS
red-shifted reporters of vesicle fusion have also been generated,
including VGLUT-2XmOr2 (Li et al., 2011), synaptobrevin 2- Generation of syn::sypHy-RGECO
mOrange (Ramirez et al., 2012), and sypHTomato (Li and Tsien, Plasmid
2012), again offering the possibility of combined imaging with SypHy was amplified by PCR using primers containing HindIII
other spectrally distinct reporters. and NotI restriction sites from CMV::sypHy A4 (a gift from
Despite the development of these tools, relatively few studies Leon Lagnado, Addgene plasmid # 24478). The resulting PCR
have been carried out in which both presynaptic calcium influx fragment was subcloned in to a modified version of the
and vesicle exocytosis have been imaged in the same small, en- pEGFP-N2 plasmid (Clontech Laboratories, USA), in which the
passant presynaptic terminals. Ermolyuk et al. (2012) used FM CMV promoter had been replaced with the human synapsin I
dyes to monitor vesicle exocytosis in conjunction with a calcium promoter. To amplify RGECO from SyRGECO (Walker et al.,
indicator introduced via a patch pipette, to examine the major 2013), a forward primer containing an XmaI site followed by a
parameters of presynaptic function including the size of the RRP, 5 amino acid linker (GSGGT) was used in conjunction with a
the probability of single vesicle fusion, calcium influx and the reverse primer containing a NotI site. When subcloned into these
sites in the syn::sypHy-EGFP-N2 plasmid the EGFP was replaced were imaged in HBS pH5.5 and HBS pH7.4 ±50 mM NH4 Cl
to generate syn::sypHy-RGECO. and the surface fraction calculated as (GpH7.4 –GpH5.5 )/(Gmax –
GpH5.5 ). Synapses in which the surface fraction was calculated to
Hippocampal Neuronal Culture and be >100% were not included (<1% of synapses).
Transfection Analysis of Live Images
Dissociated hippocampal neurons were prepared from Sprague– Images were analyzed using custom written Matlab codes
Dawley rats at embryonic day 18. Dissected hippocampi were (Mathworks, USA). Regions of interest (ROIs, 6 × 6 pixels)
treated with 5 mg/ml trypsin (Worthington, UK) for 5 min at 37 were selected for each puncta of sypHy fluorescence. Mean
and triturated with fire-polished Pasteur pipettes. Neurons were background-subtracted fluorescence intensity values were
plated on 18mm glass coverslips (Hecht Assistent, Germany) or calculated for each ROI in both sypHy (G) and RGECO (R)
35 mm Grid500 µ-dishes (Ibidi, Germany) coated with poly-D- channels. Traces were smoothed by averaging over a sliding
lysine (50 µg/ml) and laminin (20 µg/ml, both Sigma–Aldrich, window of 4 frames. Baseline fluorescence (G0 and R0 ) was
UK). Cultures media were maintained in neurobasal media measured as the average of ten frames prior to the stimulus.
supplemented with 2% B27, 2% fetal bovine serum, 1% glutamax 1G and 1R values were calculated by the change in signal
(all ThemoFisher Scientific, UK) and 1 % penicillin/streptomycin intensity from the baseline, with the peak responses defined as
(Sigma), at 37◦ C in a humidified incubator with 5% CO2 . After the maximum 1G and 1R within 40 frames of the stimulus.
3 days in vitro (DIV) the media was exchanged for serum-free Puncta in which the 1R response to the first 10 AP 20 Hz
media. At 7DIV neurons were transfected with sypHy-RGECO stimulus was greater than three times the standard deviation
using Effectene (QIAGEN, UK). After transfection neurons of the baseline were considered responding synapses and were
were maintained in serum-free media without antibiotics. Only analyzed in further images or stimulation conditions. 1G and
schedule 1 procedures performed by a competent individual 1R values for each synapse were averaged across trials and any
were used in these studies, which are exempt under the Animals synapse in which the mean peak response did not reach 3 SDs of
(Scientific Procedures) Act 1986. the mean baseline in either channel were also discarded. Before
pooling, mean 1G and 1R responses were normalized to either
Imaging of sypHy-RGECO R0 or Rmax , defined as the average of 10 images in 200 µM
Neurons were imaged at 17–21DIV in HEPES buffered saline ionomycin.
(HBS; 139 mM NaCl, 2.5 mM KCl, 10 mM HEPES, 10 mM
D-Glucose, 2 mM CaCl2 , 1.3 mM MgCl2 ; pH 7.3, 290 mOsmol) Immunofluorescence
with 2,3-dioxo-6-nitro-1,2,3,4-tetrahydrobenzo[f ]quinoxaline- After fixation cells were permeabilized with 0.2% Triton X-100 in
7-sulfonamide (NBQX), 0.025 mM amino-5-phosphonovaleric PBS for 5 min, washed three times in PBS and placed in blocking
acid (APV) and 6-imino-3-(4-methoxyphenyl)-1(6H)-pyrida- solution (3% BSA in PBS + 0.05% NaN3 ) for 60 min at room
zinebutanoic acid hydrobromide (Gabazine) (all Tocris, UK). temperature (RT), or for longer at 4◦ C. Cells were incubated in
Coverslips were placed in a field stimulation chamber containing primary antibodies diluted in blocking solution for 60 min at RT
platinum electrodes (RC49-MFS, Warner Instruments, USA). then washed five times in PBS. Secondary antibodies in blocking
For Ibidi dishes a stimulation insert (RC-37WS, Warner solution were applied for 60 min at RT and cells were washed five
Instruments) was placed inside the dish. Neurons were imaged times in PBS. Cells were mounted using Ibidi mounting medium.
on an inverted Olympus IX71 microscope, equipped with Cells were stained with chicken anti-GFP (1:1000, Abcam,
a 60x/1.42 NA oil objective. Imaging was performed with a UK) to amplify the sypHy signal, and either rabbit anti-RIM1/2
dual band-pass filter set optimized for EGFP and mCherry (1:500, Synaptic Systems, Germany) or mouse anti Synaptogmin-
(Chroma, cat number 59022) and with LED excitation light 1 (1:100, Synaptic Systems). Appropriate secondary antibodies
sources of 470 and 585 nm (CoolLED, UK) for pHluorin and tagged with Alexa 488 or 647 (ThermoFisher Scientific) were used
RGECO fluorophores, respectively. Time-lapse images were at 1:1000.
acquired at approximately 9.4 Hz using an Evolve 512 EMCCD
camera (Photometrics, USA) controlled by Slidebook software Fixed Cell Imaging and Analysis
(Intelligent Imaging Innovations, USA). Each frame consisted of The regions previously imaged live were located using the grid
a 20 ms exposure of the 470 and 585 nm LEDs sequentially, at coordinates and imaged on an inverted Nikon Eclipse Ti spectral
50 and 70% power, respectively, with 2 × 2 binning. Stimulation confocal microscope equipped with a 60x/1.40 NA oil objective
consisted of 1 ms 80 V pulses, which approximate single APs using NIS Elements software. Excitation wavelengths were 405,
(Zhao C. et al., 2011), delivered by an SD9 Stimulator (Grass 488, and 636 nm with bandpass emission filters set to 425–
Technologies, USA), controlled by Slidebook. At the end of 475 nm, 500–550 nm, and 662–737 nm. Microscope settings were
the imaging session 200 µM ionomycin was applied (Cayman kept the same for all images of the same antibodies. Z-stacks
Chemicals, USA) in either standard HBS, or HBS in which were generated from 0.15 µm optical sections, and maximum
50 mM of NaCl was replaced with 50 mM NH4Cl, to maximize projections were produced in Image-J. To align the live and fixed
sypHy-RGECO fluorescence. Neurons in Ibidi dishes were fixed images landmarks were selected in the GFP channel of both
in 4% PFA + 1% sucrose after imaging and were not treated with images, followed by affine transformation of the fixed image.
ionomycin. For calculations of surface fluorescence, synapses ROIs from the live image were scaled and overlaid on the
registered GFP image for manual confirmation of the position. (Supplementary Figure S1B). A similar correlation was observed
The GFP image was then thresholded using the ‘Moments’ when ionomycin treatment was combined with application of
algorithm in ImageJ to produce a mask, which was applied to HBS containing 50mM NH4 Cl to alkalinize all compartments
the other channels of the fixed image to avoid the inclusion of and maximize sypHy fluorescence (Supplementary Figure S1C).
neighboring puncta in the ROI. Intensity values for each ROI This contrasts with the less tight correlations observed between
were then calculated from the masked images. the baseline fluorescence values of the two channels, G0 and
R0 (Supplementary Figure S1D), and between G0 and Gmax
Statistical Analysis (Supplementary Figure S1E), which are likely affected by
Mean responses are displayed ±SEM unless otherwise stated, heterogeneity in the baseline fluorescence of sypHy. To directly
and n = the total number of synapses. As data are not normally measure the surface fraction of sypHy-RGECO, we compared the
distributed, non-parametric statistical tests have been used. difference in sypHy fluorescence in external solutions at pH5.5
and pH7.4 to the sypHy fluorescence maximized by application
of 50mM NH4 Cl. The surface fraction varied between synapses
RESULTS (Supplementary Figure S1F, mean = 30.8%, median = 24.8%) and
was slightly higher than that reported for ratio-sypHy (median
To generate a single molecule reporter of both presynaptic f surf = 20%, Rose et al., 2013). This may be due to the fact that
calcium influx and synaptic vesicle exocytosis, the R-GECO1 as basal RGECO fluorescence is low, it was necessary to identify
fragment was amplified from SyRGECO (Walker et al., 2013) transfected cells using basal sypHy fluorescence, potentially
and subcloned downstream of sypHy (Granseth et al., 2006) with biasing toward brighter cells with higher surface fractions.
a short linker sequence. The whole sequence was placed under SypHy-RGECO responses to a 10 AP stimulus from multiple
control of the human synapsin I promoter to drive expression cells were normalized and pooled, and a positive non-linear
in neurons. The resultant syn::sypHy-R-GECO1 plasmid will correlation between 1G/Rmax and 1R/Rmax was observed
hereafter be referred to as sypHy-RGECO (Figure 1A). With this (Figure 1E), as expected from the non-linear relationship
probe, presynaptic calcium influx is reported by an increase in between calcium and exocytosis (Dodge and Rahamimoff, 1967)
red fluorescence and exocytosis of synaptic vesicles by an increase and similar to that seen in other imaging studies (Zhao C.
in green fluorescence (Figure 1A). In dissociated hippocampal et al., 2011; Ariel et al., 2012; Ermolyuk et al., 2012). The spread
neurons transfected with sypHy-RGECO, punctate fluorescence of the responses suggested that individual synapses differed
in presynaptic boutons was observed in both green and red in the amount of neurotransmitter released for a given influx
channels (Figure 1B). A field stimulation of 10 action potentials of calcium, which can be evaluated by the ratio 1G/1R. We
(APs) at 20 Hz was applied and individual boutons responded observed a range of ratio values, the majority of which fell within
with increases in fluorescence in both channels (Figure 1C), a normal distribution, although a small number of synapses
the amplitudes of which (1G and 1R) were highly correlated displayed very high release in comparison to calcium influx, an
(Figure 1D). order of magnitude higher than the lowest ratios (Figure 1F).
In order to pool responses from multiple cells or synapses, Thus, there is heterogeneity in the calcium dependence of release
baseline fluorescence is often used to normalize reporter between synapses.
responses. However, for pHluorins this value can be dominated When using ratiometric imaging, differential bleaching of
by protein localized to the plasma membrane, which is the fluorophores could pose a potential problem. In boutons
approximately 20-fold brighter than the quenched vesicular imaged without stimulation, sypHy fluorescence decayed by
protein (Sankaranarayanan et al., 2000), and is highly variable only 2% during the imaging period. This decay could be fit
between synapses. One method that has been used to overcome with a double exponential, with time constants of 1.64 s and
this problem involved fusion of the red fluorescent protein 3172 s (Supplementary Figure S2A). Frames within the first
tdimer2 to the C-terminus of sypHy, generating a reporter time constant were therefore disregarded and frames from 1 s
termed ratio-sypHy (Rose et al., 2013). The fixed stoichiometry before the stimulus were used to calculate baseline fluorescence.
of the two fluorophores in this probe allowed normalization of The slow phase of bleaching, which accounted for 98.5% of the
the sypHy signals to the tdimer2 fluorescence. Using the same decay, had a negligible effect on either baseline fluorescence or
reasoning, RGECO can be used not only as a calcium indicator, maximum 1G amplitude and was not corrected for, although
but to normalize both 1G and 1R responses. As baseline this would need to be considered if imaging over a long period
RGECO fluorescence (R0 ) is relatively dim, we applied ionomycin or using this probe to measure vesicular endocytosis rates.
in 2mM extracellular calcium to maximize the RGECO signal RGECO did not exhibit bleaching, instead showing an initial
within the bouton (Rmax ). Both values (R0 and Rmax ) were highly rapid increase in fluorescence that plateaued, the reasons for
correlated (Supplementary Figure S1A) and were subsequently which will be discussed in more detail below (Supplementary
used as a normalizing factor for sypHy and RGECO responses. Figure S2B). The rapid initial changes in fluorescence suggested
After ionomycin treatment, sypHy fluorescence also gradually that sequential images may be affected by differential bleaching.
increased to a plateau presumably due to vesicle release caused To test this, cells were stimulated with 10 AP stimuli five times,
by the large influx of calcium. The maximum fluorescence in with a 1 min interval between images. The average 1G and 1R
green and red channels was also highly correlated, reflecting responses both decreased over the trials, with a slightly greater
the fixed ratio of the two fluorophores at all synapses decrease in the calcium response (Supplementary Figure S2C).
FIGURE 1 | Generation and characterization of sypHy-RGECO. (A) Schematic of the sypHy-RGECO construct and its function. Presynaptic calcium influx is
reported by an increase in RGECO fluorescence, situated on the cytoplasmic tail of synaptophysin, whilst exocytosis is reported by an increase in pHluorin
fluorescence situated in the vesicle lumen. (B) Representative images of a dissociated hippocampal cell expressing sypHy-RGECO. Scale bar 20 µm. Inset shows
merge of the two channels for the boxed region. (C) Changes in sypHy (i, 1G) and RGECO (ii, 1R) fluorescence in response to a field stimulus of 10 APs at 20 Hz.
(n = 65 synapses from 1 coverslip, traces representing individual synapse responses shown in green/red, mean response of all synapses shown in black, stimulation
period shown by black bar). (D) 1G and 1R responses are correlated (Spearman’s rank correlation r s = 0.766, p < 0.001). (E) Cells were stimulated five times with
a 10 AP 20 Hz stimulus. A mean response across trials was calculated and normalized to the maximum RGECO fluorescence in 200 µM ionomycin for each
synapse. 1G/Rmax and 1R/Rmax responses from multiple cells show a non-linear correlation (r s = 0.727, p < 0.0001, n = 526 synapses, from 9 coverslips and 5
independent cultures). The black line shows a fit to a power function of the form axb +c, where b = 2.18. (F) Distribution of values for the ratio 1G/1R, a measure of
the calcium dependence of release (for the same synapses as in E).
The reasons for these decreases are not clear, and may be ratio of the two responses remained relatively stable across
different for the two responses. RGECO did not exhibit bleaching trials, increasing by only 9% between the first and fifth repeat
across trials, whilst the baseline fluorescence of sypHy decreased (Supplementary Figure S2C). To avoid any inherent bias due to
by approximately 20% (Supplementary Figure S2C). It has the effect of repeated trials we interspersed different stimulation
previously been suggested that GECIs targeted to the presynaptic conditions, such as number or frequency of AP stimuli.
compartment can exhibit rundown in responses (Walker et al., It has previously been reported that R-GECO1 displays
2013), whereas the responses of synaptic pHluorins are similar photoactivation when illuminated with 488 nm light (Akerboom
across repeated trials (Burrone et al., 2006). However, the et al., 2013), which may explain the increase in red fluorescence
observed in unstimulated boutons (Supplementary Figure S2B). 1994; Gasparini et al., 2001). The contribution of different
To examine if this phenomenon affected sypHy-RGECO calcium channel subtypes to the presynaptic calcium signal varies not
responses, single channel images, illuminated with 585 nm only between synapses, but also with stimulation frequency
light only, were compared to our usual imaging conditions, (Ricoy and Frerking, 2014). We stimulated neurons expressing
alternating 470 and 585 nm light. 1R responses were larger in sypHy-RGECO with 10 APs at 5, 20, and 83 Hz, to examine
images taken with 585 nm light only (Supplementary Figure whether stimulation frequency affected the relationship between
S2C, upper trace), although when scaled for comparison, the 1G and 1R responses. Pooled 1G and 1R responses across
kinetics of the response were the same in both imaging all frequencies showed the expected non-linear relationship
conditions (Supplementary Figure S2C, lower trace). The (Figure 3D). Both calcium influx and vesicle exocytosis were
response amplitudes and baseline fluorescence in single channel significantly reduced at 83 Hz stimulation, which could be
and two channel images were highly correlated and, importantly, due to the failure of the cell to fire action potentials at
scaled linearly (Supplementary Figures S2D,E). These results high frequency, or to the inactivation of calcium channels
suggested that in our imaging conditions, fast alterations between (Figures 3D,E). However, the distributions of 1G/1R ratios were
blue and green excitation light reduces the size of the RGECO indistinguishable between stimulation frequencies (Figure 3F),
responses in a multiplicative manner, without the introduction of suggesting frequency did not alter the relationship between
non-linear artifacts. calcium influx and vesicle release.
To characterize the sypHy-RGECO response to different Synaptic boutons are heterogeneous in both structural and
stimulus strengths, stimuli ranging from 1 AP to 20 AP at functional parameters. Studies of the relationship between the
20 Hz were delivered by field stimulation. Both responses two show that bouton volume is a poor predictor of functional
increased linearly over the range of stimulus strengths tested properties such as presynaptic calcium influx and the synaptic
(Figures 2A,B), as has been previously reported for SyRGECO probability of release (Ermolyuk et al., 2012; Holderith et al.,
(Walker et al., 2013). Peak 1G and 1R responses to different 2012), whereas active zone size strongly correlates with function
numbers of action potentials also showed a strong linear (Holderith et al., 2012). Additionally, levels of the active zone
correlation (Figure 2C). The mean ratio between vesicle proteins RIM1/2 are highly correlated with active zone size
exocytosis and calcium influx did not differ between 5, 10, and (Holderith et al., 2012). To examine the relationship between
20 AP stimuli (Figure 2D). The mean ratio was significantly sypHy-RGECO responses and active zone size, neurons were
increased at lower stimulus strengths, although many synapses first imaged live, whilst stimulating with 10 APs at 20 Hz,
that responded at higher stimulus strengths fell below the and subsequently fixed and stained for RIM1/2. Live and fixed
detection threshold in the 1 AP and 2 AP stimuli, potentially images were aligned to allow the comparison of functional and
biasing toward synapses with higher release probabilities in structural data from individual identified synapses (Figure 4A
these groups. At higher numbers of APs, the longer stimulation and Supplementary Figure S4). We observed significant positive
times required would begin to overlap with the endocytosis correlations between RIM levels and both sypHy and RGECO
and reacidification of synaptic vesicles, making the amplitude responses (Figure 4Bi,ii). Whilst significant, these correlations
of sypHy responses harder to interpret. However, to check for are weaker than those found in ultrastructural studies, most
potential saturation of the probe we also delivered 40 and likely due to limitations in the alignment process and the lower
100 AP stimuli at 20 Hz. The calcium response indicated by resolution of light microscopy. Interestingly, RIM levels also
RGECO was saturated at 40 APs and displayed no further correlated significantly, albeit weakly, with the ratio 1G/1R
increase in amplitude to a 100 AP stimulus, whilst sypHy (Figure 4Biii), possibly reflecting the role of RIM, along with
responses continued to increase with longer stimulation trains RIM binding protein (RBP), in localizing calcium channels to
(Supplementary Figure S3). the active zone (Kaeser et al., 2011), vesicle docking (Han et al.,
We next tested the effect of changing the external calcium 2011) and coupling of calcium channels to release (Kaeser et al.,
concentration, recording sypHy-RGECO responses in both 2012).
0.5 mM and 2 mM calcium for the same set of synapses. As Similar correlations were also carried out with staining for
expected, both the calcium influx and level of vesicle release the putative calcium sensor for synchronous neurotransmitter
were significantly reduced in 0.5 mM calcium (Figure 3A), release, synaptotagmin-1 (Syt1) (Geppert et al., 1994; Fernandez-
although the positive correlation between the two responses was Chacon et al., 2001). We found that the levels of Syt1
maintained (Figure 3B). For synapses which responded above were positively correlated with the magnitude of the RGECO
threshold in both conditions, the distribution of 1G/1R ratios response, and showed a trend toward a positive correlation
was not significantly different between the two concentrations with neurotransmitter release that did not reach significance
(Figure 3C), although the ratios measured for 0.5 mM calcium (Figures 4Ci,ii). However, the ratio of release to calcium
showed a skewed distribution toward higher values. This influx at each synapse was not related to the level of Syt1
discrepancy is likely due to the saturation of the vesicular sensor (Figure 4Ciii). As the neurons were stimulated with a train of APs
of exocytosis for higher calcium concentrations, as reported at high frequency (20 Hz), both synchronous and asynchronous
previously (Ermolyuk et al., 2012). forms of release are expected to operate (Hagler and Goda,
Calcium influx at the presynaptic terminal is predominantly 2001), and thus the levels of other calcium sensors may be
mediated by CaV 2.1 and 2.2 channels, with some contribution more important in determining the total release that occurs in
from CaV 2.3 (Takahashi and Momiyama, 1993; Wheeler et al., response to this type of stimulus. Furthermore, approximately
FIGURE 2 | SypHy-RGECO responses to different stimulus strengths. A range of AP stimuli were delivered by field stimulation at 20 Hz. Each stimulus was
repeated twice, varying the order in which different numbers of APs were given, and a mean of the two responses was taken. (A) Mean sypHy and RGECO
responses to different numbers of APs, the black bar indicates the start of stimulation. (B) Peak 1G/Rmax (green) and 1R/Rmax responses (red) increase linearly over
the range of AP stimuli tested (colored lines indicate linear regression fit: sypHy, green, r 2 = 0.986, RGECO, red, r 2 = 0.961). (C) Correlation between peak 1G/Rmax
and 1R/Rmax responses over the range of APs tested (black line indicates linear fit, r 2 = 0.989). (D) Ratio 1G/1R for each of the stimulus strengths tested. Only the
ratios for 1 and 2 APs show a significant difference from the other stimuli (one way ANOVA,∗∗ p < 0.01,∗∗∗ p < 0.001). All plots show mean ±SEM, n = 171
synapses from 4 coverslips and two independent cultures.
20% of total Syt1 is found on the surface of the bouton, 2011), fused to the C-terminus of sypHy, a reporter of
not on vesicles (Fernandez-Alfonso et al., 2006), thus the vesicle exocytosis and endocytosis (Granseth et al., 2006).
level of Syt1 present at the bouton may not accurately reflect Combining the two reporters in one molecule offers several
the size of the vesicle pool. Together, our data is in broad advantages. First, it provides two independent readouts of
agreement with the general principle that synapses that have presynaptic function, without the need for co-expression of
larger active zones also show higher levels of calcium influx multiple probes. Second, as the probes are not separated
and neurotransmitter release. Although the overall levels of Syt1 spatially and a fixed ratio of the two fluorophores is present
at the synapse also broadly correlated with these functional at all synapses, it provides a way of normalizing sypHy
measures, they were a worse predictor of synapse function signals across synapses and cells, in a manner similar to
than RIM. that employed for ratio-sypHy (Rose et al., 2013). This is
particularly important for sypHy signals, where resting levels
of fluorescence vary quite dramatically from bouton to bouton
DISCUSSION and do not correlate directly with the amount probe at the
synapse, but rather to its surface levels. Finally, the fixed
We have generated a single molecule reporter, sypHy-RGECO, stoichiometry also allows a direct comparison of the two
for concurrent imaging of presynaptic calcium influx and fluorescence responses, thereby providing a direct measure of
synaptic vesicle recycling. This hybrid molecule consisted the calcium dependence of release within single presynaptic
of a red calcium indicator, R-GECO1 (Zhao Y. et al., boutons.
FIGURE 3 | Effects of altering stimulation conditions. (A–C) Dissociated hippocampal neurons in external solution (HBS) containing 0.5 mM calcium were
stimulated with a train of 10 APs at 20 Hz repeated five times 1 min apart. The solution was replaced with HBS containing 2 mM calcium and the stimulus set
repeated. Responses across trials were averaged for each synapse. (A) Fluorescence changes in sypHy (upper traces) and RGECO (lower traces) in 0.5 and 2 mM.
Traces show mean ± SEM for n = 131 synapses from 3 coverslips from 3 independent cultures. (B) Correlation between peak 1G/Rmax and 1R/Rmax responses
for these synapses (Spearman’s rank correlation r s = 0.873, p < 0.001, the black line indicates a fit to a power function of the form axb +c, where b = 1.30.)
(C) Distribution of 1G/1R ratios for synapses which responded above threshold at both calcium concentrations, ratios shown are averaged from the last two
0.5 mM trials (gray) and the first two 2 mM trials (black) (n = 47 synapses). The distribution is not significantly different between concentrations (Kolmogorov–Smirnov
test, p = 0.093). (D–F) Hippocampal neurons were stimulated with 10 APs at 5, 20, and 83 Hz. Each set of stimuli was repeated three times, varying the order in
which the frequencies were delivered and the mean of the trials was calculated. (D) Correlation between peak 1G/Rmax and 1R/Rmax values at 5 Hz (blue), 20 Hz
(red), and 83 Hz (green). Mean responses to each frequency are shown in the larger, brighter points (n = 134 synapses from 3 coverslips and three independent
cultures, the black line indicates a fit to a power function of the form axb +c, where b = 1.91). (E) Distribution of 1G/Rmax (left) and 1R/Rmax (right) responses for
these synapses show a decrease in the responses at 83 Hz compared to 20 Hz (Kolmogorov–Smirnov test, 1G/Rmax p = 0.0013, 1R/Rmax p < 0.001) but no
change in the response at 5 Hz (Kolmogorov–Smirnov test 1G/Rmax p = 0.537, 1R/Rmax p = 0.240). (F) The distribution of 1G/1R ratios does not change
between stimulus frequencies (Kolmogorov–Smirnov test 5 Hz p = 0.703, 83 Hz p = 0.282).
We observed a strong non-linear correlation between calcium both calcium and exocytosis signals. A possible source of this
influx and vesicle exocytosis across synapses that was maintained variation may be the repertoire of presynaptic calcium channel
in response to different stimulation conditions, in agreement with subtypes, which can vary between synapses (Reid et al., 1997;
many previous studies (Ariel et al., 2012; Ermolyuk et al., 2012). Nimmervoll et al., 2013). In support of this, individual boutons
Using the ratio of the sypHy response to the RGECO response displayed differential sensitivity to pharmacological blockade of
(1G/1R), we were also able to measure the calcium-dependence CaV 2.1 and 2.2 channels in both calcium influx measured by
of release at individual presynaptic boutons, a feature that has GCaMP3 and exocytosis measured by sypHTomato, when these
been difficult to study in these small CNS synapses. Interestingly, reporters were co-expressed in the same cell (Li and Tsien, 2012).
we observed heterogeneity in this value, suggesting that calcium Similarly, in another study in which calcium and vesicle dynamics
triggers vesicle fusion with different efficiencies at different were imaged in separate cells, pharmacological inhibition of
synapses. It is not yet clear what underlies this variation, as the CaV 2.1 or 2.2 led to variable blockade of neurotransmitter release
distribution of 1G/1R ratios did not change significantly under between individual synapses, although this did not alter the
different stimulation conditions, such as stimulation frequency relationship between calcium and exocytosis measured across
and extracellular calcium, despite changes in the amplitude of the synapse population (Ariel et al., 2012). However, neither
FIGURE 4 | Combining functional imaging with sypHy-RGECO and immunofluorescence. Neurons were stimulated five times with 10 APs at 20 Hz and the
responses averaged across trials. Cells were then fixed and stained for presynaptic markers and imaged using confocal microscopy. Live and fixed images were
registered and fluorescence intensities for individual synapses were measured. (A) Example of a live image of the green channel of sypHy-RGECO (left) and fixed
image (middle) of GFP and synaptotagmin-1 (Syt1) for the same cell. Right, magnified view of individual synapses. (B) Positive correlations between levels of RIM
and sypHy responses (i, green, p < 0.001), RGECO responses (ii, red, p < 0.001), and the 1G/1R ratio (iii, gray, p = 0.039), (n = 85 synapses from 3 dishes from
independent cultures). (C) Levels of Syt1 are not significantly correlated with sypHy responses (i, p = 0.059), or the 1G/1R ratio (iii, p = 0.323), but are correlated
with the calcium response (ii, p = 0.005) (n = 62 synapses from three dishes from independent cultures).
of these studies directly examined the relationship between An alternative possibility is that individual boutons differ
channel subtypes and the calcium dependence of release in in the levels or components of the molecular machinery that
individual boutons, which could be addressed using sypHy- links calcium influx to vesicle release. Of the markers that we
RGECO. examined using post hoc immunofluorescence, the level of the
calcium sensor synaptotagmin-1 was not related to the 1G/1R (Akerboom et al., 2013). When combined in the single molecule
ratio, although a weak correlation was observed with the level sypHy-RGECO, both reporters displayed a linear increase in
of RIM, an active zone protein with roles in the localization of the amplitude of the response across the range of 1–20 AP
calcium channels and vesicle docking (Han et al., 2011). Whilst stimuli delivered at 20 Hz, but at higher stimulus strengths the
absolute RIM levels are only weakly predictive of the ratio, it RGECO response became saturated. Whilst sypHy responses
is possible that more nuanced features such as the size and continued to increase to both 40 and 100 AP stimuli, the
shape of the active zone relative to the bouton volume, and amplitude of these responses will be affected by the decrease
positioning of elements such as calcium channels within it may in fluorescence caused by endocytosis and reacidifcation of
affect the calcium dependence of release at individual synapses. vesicles. The dynamic ranges of the two reporters are therefore
Combining functional imaging with structural imaging by well matched and suited to reporting activity up to 20 action
immunofluorescence, super-resolution, or electron microscopy potentials.
will enable investigation of these possibilities. Such methods One potential disadvantage of combining R-GECO1 with
could also be used to examine if differences in the calcium green sensors of vesicle exocytosis is its photoswitching behavior
dependence of release exist in different cell types, or to study in response to 488 nm light (Akerboom et al., 2013). Indeed,
the relationship between presynaptic function and postsynaptic we found that blue light exposure in our imaging conditions
properties at individual connections. did cause an offset in the basal RGECO fluorescence, as
Finally, it would be interesting to examine if the calcium well as a reduction in the calcium response. However, these
dependence of release is fixed for an individual synapse or is decreases scaled linearly with the original R-GECO1 response,
subject to regulation. Calcium influx and vesicle exocytosis have and both the kinetics of the calcium transient and the range
been measured individually after chronic changes in network of responses across synapses were preserved. SypHy-RGECO
activity, and the changes observed indicate that the relationship is therefore a useful addition to the expanding toolbox of
between the two is not altered at the population level following genetically encoded reporters, and will enable a detailed study of
homeostatic plasticity (Zhao C. et al., 2011). However, this is yet the relationship between calcium influx and vesicle dynamics at
to be studied simultaneously in single synapses. Furthermore, it individual synapses. Whilst we have shown proof of principle in
is unknown if short-term plasticity mechanisms involve changes hippocampal synapses in vitro, this reporter could also be used to
in the efficiency with which calcium triggers vesicle fusion. Tools study these processes in more intact systems, such as brain slices
such as sypHy-RGECO will enable optical interrogation of these and in vivo.
questions.
Several spectrally distinct genetically encoded reporters of
calcium and vesicle recycling have been developed, offering the AUTHOR CONTRIBUTIONS
opportunity to combine probes with different characteristics.
We selected to combine sypHy and R-GECO1 for several JB and RJ designed experiments, RJ performed experiments,
reasons; synaptophysin has previously been shown to tolerate analyzed data, and drafted the manuscript, JB and RJ edited and
addition of both pHluorin in the intravesicular loop with a approved the manuscript.
second fluorophore at the C-terminus (Rose et al., 2013), and
both sypHy and SyRGECO have previously been characterized
individually (Granseth et al., 2006; Walker et al., 2013). SypHy FUNDING
has a superior signal-to-noise ratio than synaptopHluorin and
whilst VGLUT-pHluorin offers further improvement in this This work was funded by an ERC consolidator grant, a Wellcome
aspect (Voglmaier et al., 2006), there is a risk that VGLUT1 Trust Investigator award and a Lister Institute Prize to JB.
overexpression may cause an increase in quantal content (Wojcik
et al., 2004; Wilson et al., 2005), and therefore alter the strength
of synaptic connections. Additionally, although VGLUT based ACKNOWLEDGMENTS
probes have been used in inhibitory neurons, it may impart
GABAergic boutons with unwanted properties resulting from We would like to thank Mideia Kotsogianni for the preparation
the overexpression of a transporter not usually found in these of dissociated hippocampal cultures, Robert E. Campbell for the
neurons. R-GECO1 is a sensitive calcium indicator with a R-GECO1 plasmid (Addgene plasmid 32444) and Leon Lagnado
wide dynamic range that exhibits a larger fluorescence change for the sypHy plasmid (Addgene plasmid 24478).
to field stimuli than the RCaMP1 sensors (Akerboom et al.,
2013; Walker et al., 2013). RCaMP2 (Inoue et al., 2015) offers
greater sensitivity than R-GECO1 but saturates at approximately SUPPLEMENTARY MATERIAL
8 APS and was therefore less suitable for combination with
sypHy, for which longer stimulus trains are typically used. The Supplementary Material for this article can be found
The 1F/F responses of R-GECO1 are also linear over a online at: http://journal.frontiersin.org/article/10.3389/fnsyn.
greater range of calcium concentrations than RCaMP1 sensors 2016.00021
REFERENCES Kaeser, P. S., Deng, L., Fan, M., and Sudhof, T. C., (2012). RIM genes differentially
contribute to organizing presynaptic release sites. Proc. Natl. Acad. Sci. U.S.A.
Akerboom, J., Carreras Calderon, N. Tian, L., Wabnig, S., Prigge, M., Tolo, J., et al. 109, 11830–11835. doi: 10.1073/pnas.1209318109
(2013). Genetically encoded calcium indicators for multi-color neural activity Kaeser, P. S., Deng, L., Wang, Y., Dulubova, I., Liu, X., Rizo, J., et al.
imaging and combination with optogenetics. Front. Mol. Neurosci. 6:2. doi: (2011). RIM proteins tether Ca2+ channels to presynaptic active zones via
10.3389/fnmol.2013.00002 a direct PDZ-domain interaction. Cell 144, 282–295. doi: 10.1016/j.cell.2010.
Akerboom, J., Chen, T. W., Wardill, T. J., Tian, L., Marvin, J. S., Mutlu, S., 12.029
et al. (2012). Optimization of a GCaMP calcium indicator for neural Li, H., Foss, S. M., Dobryy, Y. L., Park, C. K., Hires, S. A., Shaner, N. C., et al. (2011).
activity imaging. J. Neurosci. 32, 13819–13840. doi: 10.1523/JNEUROSCI.2601- Concurrent imaging of synaptic vesicle recycling and calcium dynamics. Front.
12.2012 Mol. Neurosci. 4:34. doi: 10.3389/fnmol.2011.00034
Ariel, P., Hoppa, M. B., and Ryan, T. A., (2012). Intrinsic variability in Pv, RRP size, Li, Y., and Tsien, R. W., (2012). pHTomato, a red, genetically encoded indicator
Ca(2+ ) channel repertoire, and presynaptic potentiation in individual synaptic that enables multiplex interrogation of synaptic activity. Nat. Neurosci. 15,
boutons. Front. Synaptic Neurosci. 4:9. doi: 10.3389/fnsyn.2012.00009 1047–1053. doi: 10.1038/nn.3126
Broussard, G. J., Liang, R., and Tian, L. (2014). Monitoring activity in neural Miesenbock, G., De Angelis, D. A., and Rothman, J. E. (1998). Visualizing secretion
circuits with genetically encoded indicators. Front. Mol. Neurosci. 7:97. doi: and synaptic transmission with pH-sensitive green fluorescent proteins. Nature
10.3389/fnmol.2014.00097 394, 192–195. doi: 10.1038/28190
Burrone, J., Li, Z., and Murthy, V. N., (2006). Studying vesicle cycling in Nakai, J., Ohkura, M., and Imoto, K. (2001). A high signal-to-noise Ca(2+ ) probe
presynaptic terminals using the genetically encoded probe synaptopHluorin. composed of a single green fluorescent protein. Nat. Biotechnol. 19, 137–141.
Nat. Protoc. 1, 2970–2978. doi: 10.1038/nprot.2006.449 doi: 10.1038/84397
Chen, T. W., Wardill, T. J., Sun, Y., Pulver, S. R., Renninger, S. L., Baohan, A., et al. Nikolaou, N., Lowe, A. S., Walker, A. S., Abbas, F., Hunter, P. R., Thompson, I. D.,
(2013). Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature et al. (2012). Parametric functional maps of visual inputs to the tectum. Neuron
499, 295–300. doi: 10.1038/nature12354 76, 317–324. doi: 10.1016/j.neuron.2012.08.040
Dana, H., Mohar, B., Sun, Y., Narayan, S., Gordus, A., Hasseman, J. P., et al. Nimmervoll, B., Flucher, B. E., and Obermair, G. J. (2013). Dominance of
(2016). Sensitive red protein calcium indicators for imaging neural activity. Elife P/Q-type calcium channels in depolarization-induced presynaptic FM dye
5:e12727. doi: 10.7554/eLife.12727 release in cultured hippocampal neurons. Neuroscience 253, 330–340. doi:
Dodge, F. A. Jr., and Rahamimoff, R. (1967). Co-operative action a calcium ions in 10.1016/j.neuroscience.2013.08.052
transmitter release at the neuromuscular junction. J. Physiol. 193, 419–432. doi: Ramirez, D. M., Khvotchev, M., Trauterman, B., and Kavalali, E. T. (2012).
10.1113/jphysiol.1967.sp008367 Vti1a identifies a vesicle pool that preferentially recycles at rest and
Dreosti, E., Odermatt, B., Dorostkar, M. M., and Lagnado, L. (2009). A genetically maintains spontaneous neurotransmission. Neuron 73, 121–134. doi:
encoded reporter of synaptic activity in vivo. Nat. Methods 6, 883–889. doi: 10.1016/j.neuron.2011.10.034
10.1038/nmeth.1399 Reid, C. A., Clements, J. D., and Bekkers, J. M. (1997). Nonuniform distribution
Ermolyuk, Y. S., Alder, F. G., Henneberger, C., Rusakov, D. A., Kullmann, of Ca2+ channel subtypes on presynaptic terminals of excitatory synapses in
D. M., and Volynski, K. E. (2012). Independent regulation of basal hippocampal cultures. J. Neurosci. 17, 2738–2745.
neurotransmitter release efficacy by variable Ca2+ influx and bouton size Ricoy, U. M., and Frerking, M. E. (2014). Distinct roles for Cav2.1-2.3 in
at small central synapses. PLoS Biol. 10:e1001396. doi: 10.1371/journal.pbio. activity-dependent synaptic dynamics. J. Neurophysiol. 111, 2404–2413. doi:
1001396 10.1152/jn.00335.2013
Fernandez-Alfonso, T., Kwan, R., and Ryan, T. A., (2006). Synaptic vesicles Rose, T., Schoenenberger, P., Jezek, K., and Oertner, T. G. (2013). Developmental
interchange their membrane proteins with a large surface reservoir during refinement of vesicle cycling at Schaffer collateral synapses. Neuron 77, 1109–
recycling. Neuron 51, 179–186. doi: 10.1016/j.neuron.2006.06.008 1121. doi: 10.1016/j.neuron.2013.01.021
Fernandez-Chacon, R., Konigstorfer, A., Gerber, S. H., Garcia, J., Matos, M. F., Sankaranarayanan, S., De Angelis, D., Rothman, J. E., and Ryan, T. A. (2000). The
Stevens, C. F., et al. (2001). Synaptotagmin I functions as a calcium regulator use of pHluorins for optical measurements of presynaptic activity. Biophys. J.
of release probability. Nature 410, 41–49. doi: 10.1038/35065004 79, 2199–2208. doi: 10.1016/S0006-3495(00)76468-X
Gasparini, S., Kasyanov, A. M., Pietrobon, D., Voronin, L. L., and Cherubini, E. Schneggenburger, R., and Neher, E. (2000). Intracellular calcium dependence of
(2001). Presynaptic R-type calcium channels contribute to fast excitatory transmitter release rates at a fast central synapse. Nature 406, 889–893. doi:
synaptic transmission in the rat hippocampus. J. Neurosci. 21, 8715–8721. 10.1038/35022702
Geppert, M., Goda, Y., Hammer, R. E., Li, C., Rosahl, T. W., Stevens, Takahashi, T., and Momiyama, A. (1993). Different types of calcium
C. F., et al. (1994). Synaptotagmin I: a major Ca2+ sensor for transmitter channels mediate central synaptic transmission. Nature 366, 156–158.
release at a central synapse. Cell 79, 717–727. doi: 10.1016/0092-8674(94) doi: 10.1038/366156a0
90556-8 Tian, L., Hires, S. A., Mao, T., Huber, D., Chiappe, M. E., Chalasani, S. H.,
Granseth, B., Odermatt, B., Royle, S. J., and Lagnado, L. (2006). Clathrin-mediated et al. (2009). Imaging neural activity in worms, flies and mice with improved
endocytosis is the dominant mechanism of vesicle retrieval at hippocampal GCaMP calcium indicators. Nat. Methods 6, 875–881. doi: 10.1038/nmeth.
synapses. Neuron 51, 773–786. doi: 10.1016/j.neuron.2006.08.029 1398
Hagler, D. J. Jr., and Goda, Y. (2001). Properties of synchronous and asynchronous Voglmaier, S. M., Kam, K., Yang, H., Fortin, D. L., Hua, Z., Nicoll, R. A., et al.
release during pulse train depression in cultured hippocampal neurons. (2006). Distinct endocytic pathways control the rate and extent of synaptic
J. Neurophysiol. 85, 2324–2334. vesicle protein recycling. Neuron 51, 71–84. doi: 10.1016/j.neuron.2006.05.027
Han, Y., Kaeser, P. S., Sudhof, T. C., and Schneggenburger, R. (2011). RIM Vyleta, N. P., and Jonas, P. (2014). Loose coupling between Ca2+ channels and
determines Ca2+ channel density and vesicle docking at the presynaptic active release sensors at a plastic hippocampal synapse. Science 343, 665–670. doi:
zone. Neuron 69, 304–316. doi: 10.1016/j.neuron.2010.12.014 10.1126/science.1244811
Heidelberger, R., and Matthews, G. (1992). Calcium influx and calcium current Walker, A. S., Burrone, J., and Meyer, M. P. (2013). Functional imaging in the
in single synaptic terminals of goldfish retinal bipolar neurons. J. Physiol. 447, zebrafish retinotectal system using RGECO. Front. Neural Circ. 7:34. doi:
235–256. doi: 10.1113/jphysiol.1992.sp019000 10.3389/fncir.2013.00034
Holderith, N., Lorincz, A., Katona, G., Rozsa, B., Kulik, A., Watanabe, M., et al. Wheeler, D. B., Randall, A., and Tsien, R. W. (1994). Roles of N-type and Q-type
(2012). Release probability of hippocampal glutamatergic terminals scales Ca2+ channels in supporting hippocampal synaptic transmission. Science 264,
with the size of the active zone. Nat. Neurosci. 15, 988–997. doi: 10.1038/ 107–111. doi: 10.1126/science.7832825
nn.3137 Wilson, N. R., Kang, J., Hueske, E. V., Leung, T., Varoqui, H., Murnick,
Inoue, M., Takeuchi, A., Horigane, S., Ohkura, M., Gengyo-Ando, K., Fujii, H., et al. J. G., et al. (2005). Presynaptic regulation of quantal size by the
(2015). Rational design of a high-affinity, fast, red calcium indicator R-CaMP2. vesicular glutamate transporter VGLUT1. J. Neurosci. 25, 6221–6234. doi:
Nat. Methods 12, 64–70. doi: 10.1038/nmeth.3185 10.1523/JNEUROSCI.3003-04.2005
Wojcik, S. M., Rhee, J. S., Herzog, E., Sigler, A., Jahn, R., Takamori, S., et al. (2004). Conflict of Interest Statement: The authors declare that the research was
An essential role for vesicular glutamate transporter 1 (VGLUT1) in postnatal conducted in the absence of any commercial or financial relationships that could
development and control of quantal size. Proc. Natl. Acad. Sci. U.S.A. 101, be construed as a potential conflict of interest.
7158–7163. doi: 10.1073/pnas.0401764101
Zhao, C., Dreosti, E., and Lagnado, L. (2011). Homeostatic synaptic plasticity Copyright © 2016 Jackson and Burrone. This is an open-access article distributed
through changes in presynaptic calcium influx. J. Neurosci. 31, 7492–7496. doi: under the terms of the Creative Commons Attribution License (CC BY). The use,
10.1523/JNEUROSCI.6636-10.2011 distribution or reproduction in other forums is permitted, provided the original
Zhao, Y., Araki, S., Wu, J., Teramoto, T., Chang, Y. F., Nakano, M., et al. (2011). author(s) or licensor are credited and that the original publication in this journal
An expanded palette of genetically encoded Ca2+ indicators. Science 333, is cited, in accordance with accepted academic practice. No use, distribution or
1888–1891. doi: 10.1126/science.1208592 reproduction is permitted which does not comply with these terms.
Flash-and-Freeze: Coordinating
Optogenetic Stimulation with Rapid
Freezing to Visualize Membrane
Dynamics at Synapses with
Millisecond Resolution
Shigeki Watanabe 1, 2*
1
Department of Cell Biology, Johns Hopkins University, Baltimore, MD, USA, 2 The Solomon H. Snyder Department of
Neuroscience, Johns Hopkins University, Baltimore, MD, USA
Direct patch-clamping of synaptic boutons shows an increase synaptic functions, but it has two major drawbacks. First, the
in the capacitance during exocytosis and a decrease during specimen must be physically attached to the electrical wire, and
endocytosis (von Gersdorff and Matthews, 1994; Sun et al., thus experiments can only be performed ex vivo. Second, only the
2002; Delvendahl et al., 2016). This method has a single vesicle surface of specimens (∼5 µm) can be frozen without ice crystal
sensitivity. However, it is blind to the locations of the membrane damage. These two factors limit its utility as a versatile tool for
insertion and removal. Furthermore, membrane trafficking studying membrane trafficking at synapses. To overcome these
within the cytosol cannot be visualized by this technique. limitations, we have developed a novel technique, flash-and-
Optical imaging methods aim to label functional pools of freeze, that combines optogenetics with high-pressure freezing
vesicles in the terminals and visualize their fates. The spatial (Watanabe et al., 2013a,b, 2014a). With this technique, non-
resolution of optical microscopy is limited to ∼200 nm by invasive light stimulation is used to induce synaptic transmission.
the diffraction of light. Because synaptic vesicles are small Specimens as thick as 200 µm can be properly frozen without ice
(∼40 nm in diameter) and hundreds of these vesicles are crystal formation. Therefore, intact animals like Caenorhabditis
clustered in a confined space, optical microscopy cannot resolve elegans (C. elegans hereafter) or an entire population of neurons
individual vesicles in the terminal. Therefore, several tricks in a dish can be studied. Here, I will describe the methods in detail
have been developed to visualize specific subsets of vesicles. and discuss its potential applications in synaptic cell biology.
For example, fluid phase markers like dextran, ferritin, and
quantum dots can be applied exogenously to label endocytosed High-Pressure Freezing
vesicles and visualize their dynamics in the terminal (Zhang An electron microscope operates under high vacuum to avoid
et al., 2007b; Park et al., 2012). Similarly, lipophilic dyes the scattering of electrons by gaseous molecules in the air. Thus,
like FM dyes and mCling can be internalized via endocytosis to observe a specimen in a transmission electron microscope,
and mark the newly reconstituted vesicles (Betz et al., 1992; it must be fixed and dehydrated. In addition, electron beam
Revelo et al., 2014). Finally, pH-sensitive fluorescent molecules must penetrate through the specimen, requiring extremely flat
such as pHluorin (Miesenböck et al., 1998) can be used to specimens with a thickness of ∼30–70 nm. For this reason,
monitor synaptic vesicle cycle. Fluorescence from pHlourin is the specimen is embedded in plastic and sectioned ultrathin.
quenched in the lumen of synaptic vesicles due to the low The sample preparation for electron microscopy often leads
pH. When exposed to the extracellular space by exocytosis, to the generation of artifacts. Fixation using aldehyde-based
the protein becomes fluorescent. The pHluorin molecules are chemicals cross-links proteins and aggregates them. Worse yet,
quenched again once they are internalized and vesicles are fully this reaction can induce fusion of synaptic vesicles (Smith and
re-acidified. These techniques allow visualization of particular Reese, 1980). Furthermore, dehydration leads to the shrinkage
vesicles (Sankaranarayanan and Ryan, 2000; Balaji and Ryan, of the membrane-bound structures and the overall changes in
2007; Armbruster et al., 2013). However, like in the capacitance the morphological architecture of the cells. Therefore, a better
measurement, it is difficult to visualize intracellular trafficking approach must be used to study membrane trafficking events at
events, limiting the interpretation of the results. synapses.
Electron microscopy, on the other hand, depicts all the One approach to avoid these artifacts is to immobilize cells
membrane-bound structures within the terminal, but only physically by rapid freezing. The freezing process, however,
captures a static image. To overcome this limitation, neurons leads to formation of ice crystals that can damage the cellular
are typically stimulated and fixed at defined time points after architecture by directly penetrating through the membrane.
the stimulation (Ceccarelli et al., 1972; Heuser and Reese, 1973; Alternatively, the solutes segregated from ice crystals can burst
Ferguson et al., 2007; Clayton et al., 2008; Matthews and Sterling, membrane due to the local changes in the osmotic pressure. To
2008; Hoopmann et al., 2010; Schikorski, 2014; Wu et al., prevent water molecules from forming ice crystals, a freezing
2014). However, the temporal resolution is limited to seconds to rate of at least 10,000 K/s must be achieved. At this rate, water
minutes for two reasons. First, to ensure the capture of endocytic molecules cannot rearrange to form ice crystals and are frozen
events, neurons are often stimulated with prolonged intense in an unordered state. The cooling rate by liquid nitrogen can
stimulation that lasts seconds to minutes. Second, diffusion exceed 16,000 K/s. Unfortunately, the heat conductance of water
and reaction of aldehyde-based chemicals are slow, requiring is poor, reducing the rate to 1000 K/s within 10 µm from the
an additional time. To visualize membrane dynamics on a point of the contact. Under high pressure (2100 bar), however,
millisecond time scale, Heuser and Reese developed the “freeze a freezing rate of 100 K/s is sufficient to freeze water in an
slammer,” which freezes tissues near instantaneously following unordered state due to the supercooling effect (Moor, 1987;
electrical stimulation (Heuser et al., 1979; Heuser and Reese, Dubochet, 2007). Thus, high-pressure freezing allows freezing of
1981). Using this device, they were able to observe vesicles fusing tissues up to 200 µm thickness or intact animals like C. elegans,
with the plasma membrane 5 ms after an electrical pulse (Heuser (∼70 µm thick) without ice crystal damage.
and Reese, 1981; Heuser et al., 1979). Although individual images
are static, membrane dynamics can be visualized in electron Optogenetics
micrographs by inducing a particular activity and freezing To visualize membrane dynamics during neurotransmission,
neurons at multiple defined time points. neurons must be stimulated and frozen at defined time points
The freeze slamming approach allows visualization of after stimulation. In the freeze slammer, specimens are physically
membrane dynamics at the temporal resolution relevant to attached to the stimulating wire, limiting the specimen choice.
Furthermore, the configuration of the high-pressure freezer feeder layer, cortices are treated with 0.05% Trypsin-EDTA for
makes it difficult to apply electrical field stimulation. To 20 min at 37◦ C. The dissociated astrocytes are then cultured in
overcome these limitations, we have coupled optogenetics with DMEM containing 10% FBS and 0.1% penicillin-streptomycin
high-pressure freezing. Channelrhodopsin is a light-sensitive for 2 weeks in a T-75 flask. Then astrocytes are transferred
cation channel isolated from Chlamydomonas reinhardtii (Nagel to the 12-well plate containing sapphire disks (5 × 104 / well)
et al., 2003). A flash of blue light opens the channel, allowing and grown for a week. Astrocyte mitosis is arrested a few
cations to flow into the cell, thereby depolarizing it. When the hours before neuron plating using fluorodeoxyuridine (80 mm).
channel is heterologously expressed in neurons, a short pulse of Media are exchanged with Neurobasal, a media containing 2%
light triggers an action potential, leading to synaptic transmission B27, 1% glutamax, and 0.1% penicillin-streptomycin, prior to
(Boyden et al., 2005; Nagel et al., 2005). Therefore, non-invasive neuron plating. Isolated hippocampi are treated with papain
stimulation can be applied to a population of neurons in a dish or (20 U/ ml) for 1 h and plated (5 × 104 /well) on top of the
intact animals. feeder layer. Neurons are allowed to grow for 2–3 weeks.
To couple optogenetic stimulation with high-pressure On DIV1-3, neurons are infected with lentivirus containing
freezing (“flash-and-freeze”), we have developed a device that Channelrhodopsin expressed from the neuron specific synapsin
interfaces with the computer of high-pressure freezer as well promoter.
as with an LED (Watanabe et al., 2013a,b, 2014a). This device
allows application of light pulses at defined time points before Caenorhabditis elegans
the specimen is frozen (see Section Materials and Methods). Transgenic animals expressing ChIEF, a variant of
Using this approach, we can visualize the membrane trafficking channelrhodopsin-2, are grown on an agar plate (50 mm)
events at synapses with a millisecond temporal resolution. seeded with 250 µl of E. coli OP50. Larval stage 4 (L4) transgenic
animals are transferred to agar plates seeded with bacteria
MATERIALS AND METHODS and 4 µl trans-retinal solution (100 mM) a day prior to the
experiment. These animals are kept in the dark before the
All experiments are performed according to the guidelines for experiments.
the animal use by the National Institute of Health. The animal
protocol is approved by the Animal Care and Use Committee Freezing
at Johns Hopkins University, School of Medicine. The graphical We have used three different high-pressure freezers: Leica EM
representations of the workflow is shown in Figure 1. The step- PACT2, HPM100, and ICE. C. elegans has been tested in the EM
by-step protocol can be found in the Supplementary information. PACT2, whereas neuronal cultures are frozen with the HPM100
or ICE. For the EM PACT2, a membrane carrier with a 100 µm-
Cell Cultures deep well is used to mount transgenic worms (Figure 2A). First,
For flash-and-freeze experiments, specimens must be mounted the well of the membrane carrier is filled with M9 worm buffer
on a substrate that is translucent like a glass coverslip but can containing 20% bovine serum albumin (BSA) as cryo-protectant.
withstand the extreme pressure application. Sapphire disks are Approximately 10–15 young adult animals are transferred into
ideal substrates, meeting these two criteria. In addition, thermal the solution. The membrane carrier is then mounted into the
conductivity of sapphire is extremely high, particularly at low modified specimen pod.
temperature, making it the perfect substrate. For culturing cells To mount cultured neuronal cells, the following procedures
on sapphire disks, the following procedures are performed. are carried out (Figure 2B). First, a sapphire disk with neurons
First, we sputter carbon on one side of 6 mm sapphire disks so is transferred into pre-warmed (37◦ C) recording solution
that we can distinguish the surface that the cells are cultured containing 140 mM NaCl, 2.4 mM KCl, 10 mM HEPES (pH
on. Then, we scratch out a letter “4” on the carbon-coated 7.5), 10 mM glucose, 4 mM CaCl2 , 1 mM MgCl2 , 3 µM
surface with a diamond scribe. Alternatively, a finder grid can NBQX, and 30 µM bicuculline. The addition of NBQX and
be placed on a sapphire disk as a mask. The sapphire disks bicuculline blocks the recurrent network activity following
are baked at 120◦ C overnight. After a brief dip in ethanol, channelrhodopsin stimulation. The sapphire disk is mounted
two sapphire disks are placed in each well of a 12-well plate on the countersink of a 6-mm middle plate. A spacer ring
and air-dried in a biosafety cabinet. Poly-D-Lysine solution (100 µm thickness) is placed on the sapphire disk so that
(acetic acid 3 parts, rat tail collagen 1 part, and poly-D-lysine cells are not crushed when the assembly is capped with
1 part) is applied directly on each sapphire disk. After 5 min, another sapphire disk. To cap the assembly, a sapphire disk
the excess solution is removed, and the plates are dried under is first dipped in the recording solution so that one side
the laminar flow. Prior to use, the 12-well plates containing of the disk carries a drop of solution. This sapphire disk
sapphire disks are sterilized by exposure to UV light for is then placed on top of the spacer ring gently so that air
20 min. bubbles are not trapped in between the disks. Two more
For hippocampal cultures, we first culture astrocytes as a spacer rings (200 µm each) are then placed to fill up the
feeder layer and then plate the neurons on top of the astrocytes. remaining space. The excess solution should be removed
Mouse brains are dissected out from newborn C57/BL6J mice by gently tapping with a small piece of filter paper. The
immediately after decapitation. Cortices and hippocampi are middle plate assembly is then sandwiched between two
isolated from each brain, in cold HBSS solution. For the astrocyte transparent plastic half cylinders. Note that light is shone
FIGURE 1 | Schematic drawings showing the experimental procedures. Specimens are mounted in the appropriate carrier and frozen immediately following
light stimulation. The freeze-substition is carried out in an automated freeze-substitution unit. Once the freeze-substitution is complete, specimens are embedded in
resin and cured for 48 h. The region of interest (the dotted line) is located using a stereoscope. The blue line indicates the direction of sectioning. The ultrathin sections
are mounted on a grid and imaged on a transmission electron microscope.
from the bottom side of the specimen assembly in the Programming Light Stimulation
HPM100 while it is from the top in the ICE. The video The modifications we have made for the EM PACT2 are described
demonstrating sample loading can be found on the Leica in detail in previous publications (Watanabe et al., 2013a,b,
website (http://www.leica-microsystems.com/science-lab/video- 2014a). For controlling the light stimulation and freezing, we
tutorials-filling-and-assembling-of-different-carriers-for-high- constructed an Arudino Uno based device that sends out 5V
pressure-freezing/). TTL signals to the LED and the high-pressure freezer at the
FIGURE 2 | Schematic drawings showing sample loading in the EM PACT2 (A) and in the HPM 100 or the EM ICE (B). For the EM PACT2, the membrane
carrier must be sealed with a sapphire anvil before the experiments. Typically, the liquid is slightly overfilled so that no air bubbles are trapped when sealing with the
sapphire anvil. For the HPM 100 and the EM ICE, liquid should be carried with the top sapphire disk to seal the assembly with no bubbles. For the EM PACT2, the
pressure is applied from the bottom of the membrane carrier while liquid nitrogen is applied from the side of the specimen. For the HPM 100 and the EM ICE,
pressurized liquid nitrogen is applied to the specimen from both sides.
desired time point. This device is driven by custom firmware applied to the specimen, the specimen is frozen to −20◦ C in 8 ms,
to produce any pattern of light pulses (i.e., single stimulus, 10 according to standard thermodynamic equations. To account for
Hz stimulation) and send out the “start” signal that triggers the variability in timing, we calculate the actual time interval post-
the freezing process (this device can be purchased from Marine hoc based on the pressure peak recorded by the accelerometer.
Reef International). Once the signal is sent, the freezer applies We have found the actual time interval to typically be within
hydraulic fluid to the specimen and pressurize the specimen. 33.6 ± 4.6 ms of the intended time.
When the pressure reaches 2000 bar, the liquid nitrogen is To introduce light stimulation capability in the HPM100,
immediately applied to the specimen. The initiation of this a similar Arudino Uno based device was constructed. The
freezing process needs to be timed so that the desired time operation of the HPM100 is fundamentally different from that
interval between the light stimulation and freezing is achieved. of the EM PACT2 in that pressurized liquid nitrogen is applied
To monitor the precise timing when the pressure is applied to to the specimen instead of a hydraulic fluid. After sending the
the specimen, we installed an accelerometer on the sample holder “start” signal, it takes 370 ms to compress the liquid nitrogen.
(also known as a “specimen bayonet”). When the pressure is The pressurized liquid nitrogen reaches the specimen in precisely
applied to the specimen, the sample holder jolts, producing a 72 ms. The specimen is frozen to −20◦ C in 8 ms from the time the
distinctive peak in the accelerometer recording. We found that it pressurized liquid nitrogen hits the specimen surface. Therefore,
requires about 170 ms for the EM PACT2 to apply the pressure to a total of 450 ms delay is expected. To monitor the precise timing,
the specimen after sending the ‘start’ signal. After the pressure is light stimulation device record the internal signal of the machine
that opens the liquid nitrogen valve in each shot. This signal using a glass knife. We collect ribbons of 250 sections from each
initiates almost invariably at 370 ms after the start signal. The animal. For mouse hippocampal neurons, we first remove the
actual timing of the freeze is adjusted post-hoc based on the valve sapphire disks from the sample by submerging them in the liquid
opening signal. nitrogen for a few seconds. About 40 sections are cut from the
For EM ICE, Leica microsystems has integrated the light exposed surface (Figure 1, bottom) and collected. These sections
stimulation control into the freezer. The machine is capable of are stained with 2.5% uranyl acetate in 70% methanol for 4 min
freezing specimen at the desired timing after the light stimulation prior to imaging.
with 1–2 ms variability.
Imaging
Freeze-Substitution A transmission electron microscope is operated at 80–120 keV.
Following high-pressure freezing, vitrified water from the Images are acquired with a digital camera. Roughly 200–300
specimen must be substituted with organic solvent to avoid images are collected from each time point.
crystallization of vitrified water as the specimen warms up to
room temperature for further processing. The freeze-substitution Image Analysis
is carried out in cryotubes containing fixatives (1% osmium The morphological analysis is performed in ImageJ using a
tetroxide, 1% glutaraldehyde, 1% milliQ water in anhydrous custom-written macro. This macro records X- and Y-coordinates
acetone) using an automated freeze-substitution unit (Leica of hand-traced membrane structures such as the plasma
AFS2). The cryotubes containing fixatives are stored under liquid membrane, active zone, membrane invaginations, synaptic
nitrogen to avoid the cross-reaction between osmium tetroxide vesicles, dense-core vesicles, large vesicles, and endosomes.
and glutaraldehyde. The sapphire disks containing the cells The positional information is first exported as text files
typically remain associated with the middle plate. To release the and then imported into MATLAB (MathWorks). We have
sapphire disks, the middle plate is quickly transferred to a cup written scripts in MATLAB to analyze the coordinates and
containing acetone, which is precooled to −90◦ C in AFS. Once calculate several statistics: distribution of vesicles from the
the middle plate reaches −90◦ C in the acetone, the sapphire plasma membrane, distribution of vesicles from the active zone,
disk often dissociates from the middle plate spontaneously. If diameter of vesicles, and number of vesicles. A normality
not, it can be released by a gentle tap. The sapphire disk is then test is performed on these data, and P-values are calculated
transferred into the cryotube containing the fixatives at −90◦ C. using student’s T-test for normally distributed data and Mann–
For C. elegans, the frozen specimens can be transferred under Whitney U-test for skewed data. The confidence level is set at
liquid nitrogen into the cryotubes containing fixatives. The 95%. For multiple comparisons, the Bonferroni correction is
tubes should be transferred quickly into the AFS. The following applied.
program is used for the freeze-substitution: −90◦ C for 5–30 h,
5◦ C/h to −20◦ C, 12 h at −20◦ C, 10◦ C/h to 20◦ C. The cryotubes RESULTS
are agitated at least twice a day during the substitution process.
Accuracy of Timing
Plastic Embedding To test the accuracy of the timing for light stimulation, we applied
Once the freeze-substitution is complete, specimens are a single stimulus and froze tissues at desired time points. We
embedded in Epon/Araldite resin. The resin is as composed of calculated the actual time at which the specimens were frozen
4.4 g Araldite 502, 6.2 g Eponate 12, 12.2 g dodecenyl succinic by monitoring the time when the pressure was applied to the
anhydrite (DDSA), and 0.8 ml benzyldimethylamine (BDMA). specimens. The pressure application can be monitored using an
First, the fixative solution is carefully removed from the cryotube, accelerometer for EM PACT 2 (Watanabe et al., 2013b, 2014a).
and specimens are washed with acetone a few times. Specimens For HPM 100, we recorded the signal that triggers opening
are then treated with 0.1% uranyl acetate in acetone for 1 h of the valve that allows pressurized liquid nitrogen to pass
to enhance the membrane contrast. Following several acetone through because specimens are frozen exactly 80 ms after the
washes, infiltration is carried out in the same cryotubes on an valve opening (see Section Materials and Methods; Watanabe
orbital shaker: 30% for 2–5 h, 70% for 3–6 h, and 90% overnight et al., 2013a). On the EM ICE, we can measure the interval
at room temperature. The next day, specimens are transferred between the end of the desired stimulation program and the
into the caps of BEEM capsules containing 100% Epon/Araldite time at which the sensor, which is located close to the specimen,
resin. The resin is replaced three times (2 h each). The specimens reaches 0◦ C. The sensor is placed before the specimen, and it
are cured in an oven (60◦ C) for 48 h. requires additional 3-6 ms for the specimen to experience the
same magnitude of cooling.
Sectioning We recorded the values from each machine on at least three
For observation in a transmission electron microscope, different experimental days. Out of 52 shots taken on the EM
specimens must be sliced thin so that electrons can penetrate PACT2 (4 experiments, 13 shots/experiment), the desired time
through the tissue and generate an image. Ultrathin sections point was achieved in only 5 shots (Figure 3). The actual timing
(40 nm) of specimens are cut using a diamond knife and collected was variable, ranging from −102 to 110 ms. Because of this
on pioloform-coated single slot grids. For C. elegans, we orient variability, the actual timing is always calculated post-hoc for each
the animal so that it is perpendicular to the sectioning surface shot on the EM PACT2. On the HPM100, nearly two thirds of
(Figure 1, bottom). We trim the animal to the reflex of the gonad the shots (57/90) achieved the desired time point (3 experiments,
FIGURE 3 | A bar graph showing the time difference between the DISCUSSION
desired time point and the actual time point at which specimens are
frozen. The variation is largest in the EM PACT2, with the average delay of Comparisons among Three Instruments
33.6 ± 4.6 ms (N = 4 experiments; n = 52 shots). The average delay of the To visualize membrane dynamics at synapses, we developed
HPM100 is 1.9 ± 0.3 ms (N = 3 experiments; n = 90 shots) while that of the
EM ICE is 2.2 ± 0.1 (N = 3 experiments; n = 72 shots).
a technique, “flash-and-freeze,” which couples optogenetic
stimulation of neurons with high-pressure freezing. Here, we
FIGURE 4 | Representative micrographs showing exocytic omega figures from the flash-and-freeze experiments. (A) C. elegans neuromuscular junctions,
stimulated for 20 ms, and frozen at the end of the light pulse. Mouse hippocampal neurons, stimulated for 10 ms, frozen 5 ms after the end of the pulse using HPM
100 (B), and EM ICE (C). These structures are rarely observed in non-stimulated control or in specimens frozen at later time points after the stimulation.
compared three instruments that allow such experiments: EM that take place during these manipulations can be preserved
PACT2, HPM100, and EM ICE. There are advantages and through the flash-and-freeze approach and visualized with a
disadvantages for each instrument. The major advantage of the millisecond temporal resolution. Furthermore, there are neurons
EM PACT 2 is in its potential application - higher intensity like photoreceptors that are naturally sensitive to light. The
of ultraviolet light can be applied to specimens because light flash-and-freeze approach can be readily adapted to address the
only has to penetrate through a clear sapphire anvil. Most caged mechanism underlying tonic release of neurotransmitters from
compounds are uncaged with ultraviolet light. In both HPM100 these terminals and how exocytosis is mediated by synaptic
and EM ICE, light must travel through an optical fiber and ribbons, electron dense proteinaceous protrusion in the center of
penetrate through a plastic sample holder that absorbs ultraviolet the active zone (LoGiudice and Matthews, 2009).
light. Thus, experiments using caged compound are more feasible
in the EM PACT2. However, the EM PACT2 required many Potential Difficulties
trials before the desired time point was reached. Besides the Membrane dynamics are inferred by observing the membrane
unreliability, the major disadvantage of the EM PACT2 is the morphology at defined time points, but individual electron
size of the specimen cup. A specimen cup has a dimension of 1.6 micrographs still capture static images of cells. The flash-and-
mm diameter × 100 or 200 µm depth. Cells must be cultured on freeze technique cannot follow the membrane dynamics in a
sapphire disks with a diameter of 1.4 mm. These small sapphire single cell. Therefore, to reconstruct the membrane dynamics,
disks float in culture media, making it difficult to culture cells. hundreds of images must be acquired and analyzed from each
On the other hand, sapphire disks as large as 6 mm can be time point. For membrane trafficking at synapses, an even
loaded into the HPM100 or the EM ICE. Both the HPM100 and higher number may be required due to the heterogeneity in the
the EM ICE are essentially comparable in terms of the freezing release probability (Rosenmund and Stevens, 1996)—not every
quality and the temporal precision of light stimulation. However, action potential leads to fusion of synaptic vesicles. We have
HPM100 is more flexible in terms of its potential applications, estimated that about 20–30% of synapses are activated with a
as the light stimulation device and programs are customizable. single light pulse in mouse hippocampal neurons (Watanabe
With the expansion in the repertoire of optogenetic tools or light- et al., 2013a, 2014b). To produce data with statistical significance,
sensitive compounds, cellular activity can be controlled using we have analyzed over 10,000 images. These analyses were
different wavelengths of light, and it is possible to develop a performed blind to genotypes and time points, and thus the
multi-color system for the HPM100. The major disadvantages average frequency of morphological features at a particular
of HPM 100 are that it requires an external device and that time point can be determined from these images. To expedite
the actual freezing point must be calculated based on when analysis and increase the statistical power, automation is in an
the liquid nitrogen valve opens each time. On the contrary, immediate need. However, with the recent advances in methods
EM ICE does not require a custom device or an external light for automated image acquisition and analysis (Potter et al., 1999;
source—everything is integrated. Furthermore, it achieves the Denk and Horstmann, 2004; Knott et al., 2008; Hayworth et al.,
most precise control of the light stimulation and the fastest 2014), achieving such a goal is likely within reach.
freezing rate, requiring only 3 ms from the initial temperature
drop to −20◦ C. Despite these differences among instruments, AUTHOR CONTRIBUTIONS
exocytosis of synaptic vesicles, which occurs on a millisecond
time scale, were captured in all three instruments, suggesting SW performed all the experiments, analyzed the data, and wrote
that any of these machines are compatible with flash-and-freeze the manuscript.
experiments. Therefore, the choice should be made depending on
the potential applications. ACKNOWLEDGMENTS
Potential Applications I would like to thank Erik Jorgensen, Christian Rosenmund, and
The flash-and-freeze approach can be applied to study many all the members in their laboratories for the help in performing
cell biological problems. We have been using channelrhodopsin the original experiments. The light stimulation device for Leica
to stimulate neuronal activity and capture membrane dynamics EM PACT 2 and HPM 100 is designed and developed by M.
at synapses. Neuronal activity or cellular functions can be Wayne Davis. I would like to thank the Grass Foundation and
manipulated using light-sensitive molecules such as bacterial the Marine Biological Laboratory at Woods Hole for space,
opsins (Han and Boyden, 2007; Zhang et al., 2007a; Berthold equipment and funding for performing these experiments. I
et al., 2008; Chow et al., 2010), caged compounds (Walker et al., would like to thank Sumana Raychaudhuri and Edward Hujber
1986; Tsien et al., 1987; Milburn et al., 1989; Adams and Tsien, for a critical reading of the manuscript. The research is supported
1993; Wieboldt et al., 1994), light-sensitive proteins (Levskaya by Johns Hopkins University startup fund.
et al., 2009; Wu et al., 2009; Yazawa et al., 2009; Idevall-Hagren
et al., 2012; Zhou et al., 2012; Guntas et al., 2015), and photo- SUPPLEMENTARY MATERIAL
isomerizable molecule (Banghart et al., 2004; Kramer et al., 2009).
These tools can be used to inhibit or activate neuronal activity, The Supplementary Material for this article can be found
protein function, lipid composition, signaling cascades, and ion online at: http://journal.frontiersin.org/article/10.3389/fnsyn.
composition in vesicular structures. The morphological changes 2016.00024
REFERENCES Heuser, J. E., and Reese, T. S. (1981). Structural changes after transmitter
release at the frog neuromuscular junction. J. Cell Biol. 88, 564–580. doi:
Adams, S. R., and Tsien, R. Y. (1993). Controlling cell chemistry with 10.1083/jcb.88.3.564
caged compounds. Annu. Rev. Physiol. 55, 755–784. doi: 10.1146/annurev. Heuser, J. E., Reese, T. S., Dennis, M. J., Jan, Y., Jan, L., and Evans, L. (1979).
ph.55.030193.003543 Synaptic vesicle exocytosis captured by quick freezing and correlated with
Armbruster, M., Messa, M., Ferguson, S. M., De Camilli, P., and Ryan, T. A. (2013). quantal transmitter release. J. Cell Biol. 81, 275–300. doi: 10.1083/jcb.81.2.275
Dynamin phosphorylation controls optimization of endocytosis for brief action Hoopmann, P., Punge, A., Barysch, S. V., Westphal, V., Bückers, J., Opazo, F., et al.
potential bursts. Elife 2:e00845. doi: 10.7554/eLife.00845 (2010). Endosomal sorting of readily releasable synaptic vesicles. Proc. Natl.
Balaji, J., and Ryan, T. A. (2007). Single-vesicle imaging reveals that synaptic vesicle Acad. Sci. U.S.A. 107, 19055–19060. doi: 10.1073/pnas.1007037107
exocytosis and endocytosis are coupled by a single stochastic mode. Proc. Natl. Idevall-Hagren, O., Dickson, E. J., Hille, B., Toomre, D. K., and De Camilli, P.
Acad. Sci. U.S.A. 104, 20576–20581. doi: 10.1073/pnas.0707574105 (2012). Optogenetic control of phosphoinositide metabolism. Proc. Natl. Acad.
Banghart, M., Borges, K., Isacoff, E., Trauner, D., and Kramer, R. H. (2004). Light- Sci. U.S.A. 109, E2316–E2323. doi: 10.1073/pnas.1211305109
activated ion channels for remote control of neuronal firing. Nat. Neurosci. 7, Knott, G., Marchman, H., Wall, D., and Lich, B. (2008). Serial section scanning
1381–1386. doi: 10.1038/nn1356 electron microscopy of adult brain tissue using focused ion beam milling. J.
Berthold, P., Tsunoda, S. P., Ernst, O. P., Mages, W., Gradmann, D., Neurosci. 28, 2959–2964. doi: 10.1523/JNEUROSCI.3189-07.2008
and Hegemann, P. (2008). Channelrhodopsin-1 initiates phototaxis and Kramer, R. H., Fortin, D. L., and Trauner, D. (2009). New photochemical tools
photophobic responses in chlamydomonas by immediate light-induced for controlling neuronal activity. Curr. Opin. Neurobiol. 19, 544–552. doi:
depolarization. Plant Cell 20, 1665–1677. doi: 10.1105/tpc.108.057919 10.1016/j.conb.2009.09.004
Betz, W. J., Mao, F., and Bewick, G. S. (1992). Activity-dependent fluorescent Levskaya, A., Weiner, O. D., Lim, W. A., and Voigt, C. A. (2009). Spatiotemporal
staining and destaining of living vertebrate motor nerve terminals. J. Neurosci. control of cell signalling using a light-switchable protein interaction. Nature
12, 363–375. 461, 997–1001. doi: 10.1038/nature08446
Boyden, E. S., Zhang, F., Bamberg, E., Nagel, G., and Deisseroth, K. (2005). LoGiudice, L., and Matthews, G. (2009). The role of ribbons at sensory synapses.
Millisecond-timescale, genetically targeted optical control of neural activity. Neuroscientist 15, 380–391. doi: 10.1177/1073858408331373
Nat. Neurosci. 8, 1263–1268. doi: 10.1038/nn1525 Matthews, G., and Sterling, P. (2008). Evidence that vesicles undergo
Ceccarelli, B., Hurlbut, W. P., and Mauro, A. (1972). Depletion of vesicles from compound fusion on the synaptic ribbon. J. Neurosci. 28, 5403–5411.
frog neuromuscular junctions by prolonged tetanic stimulation. J. Cell Biol. 54, doi: 10.1523/JNEUROSCI.0935-08.2008
30–38. doi: 10.1083/jcb.54.1.30 Miesenböck, G., De Angelis, D. A., and Rothman, J. E. (1998). Visualizing secretion
Chow, B. Y., Han, X., Dobry, A. S., Qian, X., Chuong, A. S., Li, M., et al. and synaptic transmission with pH-sensitive green fluorescent proteins. Nature
(2010). High-performance genetically targetable optical neural silencing 394, 192–195. doi: 10.1038/28190
by light-driven proton pumps. Nature 463, 98–102. doi: 10.1038/nature Milburn, T., Matsubara, N., Billington, A. P., Udgaonkar, J. B., Walker, J. W.,
08652 Carpenter, B. K., et al. (1989). Synthesis, photochemistry, and biological activity
Clayton, E. L., Evans, G. J. O., and Cousin, M. A. (2008). Bulk synaptic vesicle of a caged photolabile acetylcholine receptor ligand. Biochemistry 28, 49–55.
endocytosis is rapidly triggered during strong stimulation. J. Neurosci. 28, doi: 10.1021/bi00427a008
6627–6632. doi: 10.1523/JNEUROSCI.1445-08.2008 Moor, H. (1987). “Theory and practice of high pressure freezing,” in
Delvendahl, I., Vyleta, N. P., von Gersdorff, H., and Hallermann, S. (2016). Cryotechiniques in Biological Electron Microscopy, eds R. A. Steinbrecht and
Fast, temperature-sensitive and clathrin-independent endocytosis at central K. Zierold (Berlin: Springer-Verlag), 175–191.
synapses. Neuron 90, 492–498. doi: 10.1016/j.neuron.2016.03.013 Nagel, G., Brauner, M., Liewald, J. F., Adeishvili, N., Bamberg, E., and Gottschalk,
Denk, W., and Horstmann, H. (2004). Serial block-face scanning electron A. (2005). Light activation of channelrhodopsin-2 in excitable cells of
microscopy to reconstruct three-dimensional tissue nanostructure. PLoS Biol. Caenorhabditis elegans triggers rapid behavioral responses. Curr. Biol. 15,
2:e329. doi: 10.1371/journal.pbio.0020329 2279–2284. doi: 10.1016/j.cub.2005.11.032
Dittman, J., and Ryan, T. A. (2009). Molecular circuitry of endocytosis at nerve Nagel, G., Szellas, T., Huhn, W., Kateriya, S., Adeishvili, N., Berthold, P., et al.
terminals. Annu. Rev. Cell Dev. Biol. 25, 133–160. doi: 10.1146/annurev. (2003). Channelrhodopsin-2, a directly light-gated cation-selective membrane
cellbio.042308.113302 channel. Proc. Natl. Acad. Sci. U.S.A. 100, 13940–13945. doi: 10.1073/
Dubochet, J. (2007). The physics of rapid cooling and its implications for pnas.1936192100
cryoimmobilization of cells. Methods Cell Biol. 79, 7–21. doi: 10.1016/s0091- Neher, E., and Marty, A. (1982). Discrete changes of cell membrane capacitance
679x(06)79001-x observed under conditions of enhanced secretion in bovine adrenal chromaffin
Ferguson, S. M., Brasnjo, G., Hayashi, M., Wölfel, M., Collesi, C., Giovedi, S., cells. Proc. Natl. Acad. Sci.U.S.A. 79, 6712–6716. doi: 10.1073/pnas.79.21.6712
et al. (2007). A selective activity-dependent requirement for dynamin 1 in Park, H., Li, Y., and Tsien, R. W. (2012). Influence of synaptic vesicle position on
synaptic vesicle endocytosis. Science 316, 570–574. doi: 10.1126/science.11 release probability and exocytotic fusion mode. Science 335, 1362–1366. doi:
40621 10.1126/science.1216937
Guntas, G., Hallett, R. A., Zimmerman, S. P., Williams, T., Yumerefendi, H., Bear, Potter, C. S., Chu, H., Frey, B., Green, C., Kisseberth, N., Madden, T. J., et al. (1999).
J. E., et al. (2015). Engineering an improved light-induced dimer (iLID) for Leginon: a system for fully automated acquisition of 1000 electron micrographs
controlling the localization and activity of signaling proteins. Proc. Natl. Acad. a day. Ultramicroscopy 77, 153–161. doi: 10.1016/S0304-3991(99)00043-1
Sci. U.S.A. 112, 112–117. doi: 10.1073/pnas.1417910112 Revelo, N. H., Kamin, D., Truckenbrodt, S., Wong, A. B., Reuter-Jessen, K.,
Hamill, O. P., Marty, A., Neher, E., Sakmann, B., and Sigworth, F. J. (1981). Reisinger, E., et al. (2014). A new probe for super-resolution imaging of
Improved patch-clamp techniques for high-resolution current recording membranes elucidates trafficking pathways. J. Cell Biol. 205, 591–606. doi:
from cells and cell-free membrane patches. Pflügers Arch. 391, 85–100. doi: 10.1083/jcb.201402066
10.1007/BF00656997 Rosenmund, C., and Stevens, C. F. (1996). Definition of the readily releasable
Han, X., and Boyden, E. S. (2007). Multiple-color optical activation, silencing, pool of vesicles at hippocampal synapses. Neuron 16, 1197–1207. doi:
and desynchronization of neural activity, with single-spike temporal resolution. 10.1016/S0896-6273(00)80146-4
PLoS ONE 2:e299. doi: 10.1371/journal.pone.0000299 Saheki, Y., and De Camilli, P. (2012). Synaptic vesicle endocytosis. Cold Spring
Hayworth, K. J., Morgan, J. L., Schalek, R., Berger, D. R., Hildebrand, D. G. C., Harb. Perspect. Biol. 4:a005645. doi: 10.1101/cshperspect.a005645
and Lichtman, J. W. (2014). Imaging ATUM ultrathin section libraries with Sankaranarayanan, S., and Ryan, T. A. (2000). Real-time measurements of vesicle-
WaferMapper: a multi-scale approach to EM reconstruction of neural circuits. SNARE recycling in synapses of the central nervous system. Nat. Cell Biol. 2,
Front. Neural Circuits 8:68. doi: 10.3389/fncir.2014.00068 197–204. doi: 10.1038/35008615
Heuser, J. E., and Reese, T. S. (1973). Evidence for recycling of synaptic vesicle Schikorski, T. (2014). Readily releasable vesicles recycle at the active zone of
membrane during transmitter release at the frog neuromuscular junction. hippocampal synapses. Proc. Natl. Acad. Sci. U.S.A. 111, 5415–5420. doi:
J. Cell Biol. 57, 315–344. doi: 10.1083/jcb.57.2.315 10.1073/pnas.1321541111
Smith, J. E., and Reese, T. S. (1980). Use of aldehyde fixatives to determine the rate properties, and activation of glutamate receptors on a microsecond time scale.
of synaptic transmitter release. J. Exp. Biol. 89, 19–29. Proc. Natl. Acad. Sci. U.S.A. 91, 8752–8756. doi: 10.1073/pnas.91.19.8752
Sun, J.-Y., Wu, X.-S., and Wu, L.-G. (2002). Single and multiple vesicle fusion Wu, Y. I., Frey, D., Lungu, O. I., Jaehrig, A., Schlichting, I., Kuhlman, B., et al.
induce different rates of endocytosis at a central synapse. Nature 417, 555–559. (2009). A genetically encoded photoactivatable Rac controls the motility of
doi: 10.1038/417555a living cells. Nature 461, 104–108. doi: 10.1038/nature08241
Tsien, R. W., Fox, A. P., Hess, P., McCleskey, E. W., Nilius, B., Nowycky, M. C., Wu, Y., O’Toole, E. T., Girard, M., Ritter, B., Messa, M., Liu, X., et al.
et al. (1987). Multiple types of calcium channel in excitable cells. Soc. Gen. (2014). A dynamin 1-, dynamin 3- and clathrin-independent pathway of
Physiol. Ser. 41, 167–187. synaptic vesicle recycling mediated by bulk endocytosis. Elife 3:e01621. doi:
von Gersdorff, H., and Matthews, G. (1994). Dynamics of synaptic vesicle fusion 10.7554/eLife.01621
and membrane retrieval in synaptic terminals. Nature 367, 735–739. doi: Yazawa, M., Sadaghiani, A. M., Hsueh, B., and Dolmetsch, R. E. (2009). Induction
10.1038/367735a0 of protein-protein interactions in live cells using light. Nat. Biotechnol. 27,
Walker, J. W., McCray, J. A., and Hess, G. P. (1986). Photolabile protecting 941–945. doi: 10.1038/nbt.1569
groups for an acetylcholine receptor ligand. Synthesis and photochemistry of Zhang, F., Wang, L.-P., Brauner, M., Liewald, J. F., Kay, K., Watzke, N., et al.
a new class of o-nitrobenzyl derivatives and their effects on receptor function. (2007a). Multimodal fast optical interrogation of neural circuitry. Nature 446,
Biochemistry 25, 1799–1805. doi: 10.1021/bi00355a052 633–639. doi: 10.1038/nature05744
Watanabe, S., Davis, M. W., and Jorgensen, E. M. (2014a). “Flash-and-freeze Zhang, Q., Cao, Y.-Q., and Tsien, R. W. (2007b). Quantum dots provide an optical
electron microscopy: coupling optogenetics with high-pressure freezing,” in signal specific to full collapse fusion of synaptic vesicles. Proc. Natl. Acad. Sci.
Nanoscale Imaging of Synapses, eds U. V. Nägerl and A. Triller (New York, NY: U.S.A. 104, 17843–17848. doi: 10.1073/pnas.0706906104
Springer), 43–57. Zhou, X. X., Chung, H. K., Lam, A. J., and Lin, M. Z. (2012). Optical control
Watanabe, S., Liu, Q., Davis, M. W., Hollopeter, G., Thomas, N., Jorgensen, N. B., of protein activity by fluorescent protein domains. Science 338, 810–814. doi:
et al. (2013b). Ultrafast endocytosis at Caenorhabditis elegans neuromuscular 10.1126/science.1226854
junctions. Elife 2:e00723. doi: 10.7554/eLife.00723
Watanabe, S., Rost, B. R., Camacho-Pérez, M., Davis, M. W., Söhl- Conflict of Interest Statement: The author declares that the research was
Kielczynski, B., Rosenmund, C., et al. (2013a). Ultrafast endocytosis at conducted in the absence of any commercial or financial relationships that could
mouse hippocampal synapses. Nature 504, 242–247. doi: 10.1038/nature be construed as a potential conflict of interest.
12809
Watanabe, S., Trimbuch, T., Camacho-Pérez, M., Rost, B. R., Brokowski, Copyright © 2016 Watanabe. This is an open-access article distributed under the
B., Söhl-Kielczynski, B., et al. (2014b). Clathrin regenerates synaptic terms of the Creative Commons Attribution License (CC BY). The use, distribution or
vesicles from endosomes. Nature 515, 228–233. doi: 10.1038/nature reproduction in other forums is permitted, provided the original author(s) or licensor
13846 are credited and that the original publication in this journal is cited, in accordance
Wieboldt, R., Gee, K. R., Niu, L., Ramesh, D., Carpenter, B. K., and Hess, with accepted academic practice. No use, distribution or reproduction is permitted
G. P. (1994). Photolabile precursors of glutamate: synthesis, photochemical which does not comply with these terms.
The postsynaptic density (PSD), apparent in electron micrographs as a dense lamina just
beneath the postsynaptic membrane, includes a deeper layer, the “pallium”, containing
a scaffold of Shank and Homer proteins. Though poorly defined in traditionally prepared
thin-section electron micrographs, the pallium becomes denser and more conspicuous
during intense synaptic activity, due to the reversible addition of CaMKII and other
proteins. In this Perspective article, we review the significance of CaMKII-mediated
recruitment of proteins to the pallium with respect to both the trafficking of receptors
and the remodeling of spine shape that follow synaptic stimulation. We suggest that
the level and duration of CaMKII translocation and activation in the pallium will shape
activity-induced changes in the spine.
Keywords: postsynaptic density, PSD, pallium, electron microscopy, EM, Shank, Homer, CaMKII
Excitatory glutamatergic synapses examined with the electron microscope typically display a
pronounced postsynaptic density (PSD), which appears in conventional electron micrographs as
an approximately 30 nm thick electron-dense structure applied to the postsynaptic membrane
(Palay, 1956). The large majority of excitatory synapses in the vertebrate CNS release glutamate
as neurotransmitter. Ionotropic glutamate receptors concentrate at the PSD, where specialized
Edited by: molecules anchor them and regulate their trafficking; modulation of their expression and
George Augustine, trafficking plays a key role in synaptic plasticity. Decades of research show that much of the
Nanyang Technologicial University
story of this modulation resides in the PSD.
(NTU), Singapore
Using special stains, Valtschanoff and Weinberg (2001) uncovered a distinct layer ∼50 nm thick
Reviewed by: immediately adjacent and deep to the PSD. Though previously reported, this ‘‘subsynaptic web’’
Eric Hanse,
University of Gothenburg, Sweden
(DeRobertis, 1964) had been largely ignored for decades, because it is difficult to distinguish from
Anne McKinney, the rest of the spine cytoplasm after standard histological procedures. Extending this work, Petralia
McGill University, Canada et al. (2005) reported high levels of immunoEM label for both Shank and Homer in the same
*Correspondence: region, which was referred to as the ‘‘subjacent’’ area.
Ayse Dosemeci The subsynaptic web becomes more prominent after short bouts of stimulation, reflecting
[email protected] the reversible incorporation of additional proteins. Because this part of the PSD comprises
different proteins and displays a more labile organization than the ‘‘core’’ layer lining the synaptic
Received: 17 May 2016
Accepted: 25 July 2016
membrane, it merits a distinct name; we propose to designate this separate structural and functional
Published: 19 August 2016 layer of the PSD as its pallium or mantle (Figure 1). Here we present a perspective on PSD structure
Citation:
that highlights potential functions carried out in the pallium.
Dosemeci A, Weinberg RJ, Reese TS
and Tao-Cheng J-H (2016) The
Postsynaptic Density: There Is More
CORE LAYER OF THE PSD
than Meets the Eye.
Front. Synaptic Neurosci. 8:23. Though highly variable, PSDs in the rodent forebrain contain on average 300–400 copies of PSD-95
doi: 10.3389/fnsyn.2016.00023 and related membrane-associated guanylate kinase (MAGUK) molecules (Chen et al., 2005;
Names used in the present article are listed along with synonyms.
Shank (SH3 and multiple ankyrin repeat domains protein) also called ProSAP
(Proline-rich synapse-associated protein), Synamon and CortBP (Cortactin
binding protein).
FIGURE 2 | Label for two PSD scaffold proteins, Shank and Homer, lies mostly outside the PSD core. (A) Electron micrographs of asymmetric synapses in
cultured hippocampal neurons immunogold labeled with antibodies for three major PSD scaffold proteins. Silver-enhanced gold label appears as black particles of
heterogeneous size (EM in middle panel from Tao-Cheng et al., 2015). The PSD core with PSD-95 label and the PSD pallium with Shank and Homer labels are
marked by brackets on the left and middle panels respectively. Scale bar: 0.1 µm. (B) Frequency histograms depicting the distribution of label at the postsynaptic
compartment in a typical experiment. While label for PSD-95 is within the electron-dense material, label for Shank and Homer is concentrated in a deeper region we
designate as the pallium of the PSD.
with the PSD core, whereas the cytoplasmic side, characterized Role of CaMKII as a Kinase
by an extremely rough surface (Petersen et al., 2003) merges CaMKII, a Ca2+ -regulated serine-threonine protein kinase, is
imperceptibly into the central zone of the spine head. present at very high concentrations in neurons, with a relative
abundance reaching ∼1–2% of total protein in the cerebral cortex
and hippocampus (Erondu and Kennedy, 1985). Within neurons,
ACTIVITY-INDUCED CHANGES AT THE CaMKII is especially prominent in the PSD pallium (Petralia
PALLIUM et al., 2005; Ding et al., 2013). Activation of CaMKII plays
a pivotal role in certain types of synaptic modification, most
The Pallium Becomes More Electron- notably NMDA-dependent LTP (reviews: Lisman et al., 2012;
Dense as CaMKII Accumulates Shonesy et al., 2014).
Thin-section EM of cultured hippocampal neurons reveals The CaMKII holoenzyme is a dodecamer made of individually
increased electron density at the pallium under excitatory active subunits, with α- and β-isoforms prevalent in brain.
conditions (Figure 3: top panels), a phenomenon we have Subunits within a single holoenzyme can phosphorylate
described as PSD ‘‘thickening’’ (Dosemeci et al., 2001). one another, a process called autophosphorylation. Upon
The increase in electron density likely reflects the addition autophosphorylation in the presence of Ca2+ , CaMKII becomes
of proteins, and immunoEM shows significant increase in autonomous, allowing enzymatic activity to persist beyond
CaMKII immunolabel within the PSD complex under excitatory the duration of Ca2+ elevation (reviews: Hell, 2014; Shonesy
conditions (Figure 3: bottom panels, Dosemeci et al., 2001). et al., 2014). Autonomous CaMKII can remain at the PSD
Similar morphological changes at the PSD, accompanied by the pallium after the cessation of excitatory stimuli (Dosemeci et al.,
accumulation of CaMKII and other proteins, also occur under 2002; Otmakhov et al., 2004). Continued co-localization of
the excitotoxic conditions promoted by ischemia (Suzuki et al., autonomous CaMKII with its substrates at the pallium would
1994; Hu et al., 1998; Martone et al., 1999). The duration of the counteract dephosphorylation by phosphatases and thus help
morphological and compositional changes at the PSD appears maintain phosphorylation of CaMKII substrates.
to vary according to the type of stimulation; importantly, more The accumulation and activation of CaMKII at the pallium
sustained modification is observed following an LTP-inducing triggers further changes at the PSD. Some of these have been
protocol (Otmakhov et al., 2004). studied in depth, such as the incorporation of AMPA receptors
Associated 1 (AIDA-1, also known as EB-1 and ANKS1B) is a TABLE 1 | Redistribution of postsynaptic density (PSD) components
major component of the PSD complex, with a PSD-95/AIDA- during activity.
1/GKAP stoichiometry of 2:1:1 (Lowenthal et al., 2015). Like
CaMKII SynGAP AIDA-1 Shank CYLD
SynGAP, AIDA-1 binds directly to PSD-95 at the same domains
that bind TARPs (Jordan et al., 2007) and therefore could also Core decrease decrease
interfere with the anchoring of AMPA receptors. By immunoEM, Pallium increase increase increase increase increase
the AIDA-1 label is mostly located within the PSD core at rest Changes observed in immunogold label density for selected proteins following
(Jacob et al., 2010; Dosemeci et al., 2015). Under excitatory treatment of cultured hippocampal neurons with media containing NMDA or
conditions AIDA-1 label at the PSD core is significantly high K+ .
reduced with a concomitant increase at the pallium (Dosemeci
et al., 2015). The reversible CaMKII-mediated redistribution increases the density of mushroom-shaped spines; importantly,
of AIDA-1 at the PSD under excitatory conditions (Dosemeci its association with Shank is necessary for Abp1-mediated
et al., 2016) parallels that of SynGAP. We speculate that different regulation of spine morphology (Haeckel et al., 2008).
regulatory mechanisms may trigger dissociation of SynGAP and Another protein that associates with both Shanks and
AIDA-1 from PSD-95. actin is cortactin, although immunoEM shows that cortactin
concentrates mainly at the central region of the spine (Racz
and Weinberg, 2004). In contrast to Abp1, cortactin exits the
Shank Levels Increase in the Pallium spines during activity (Hering and Sheng, 2003). While Abp1
Shanks, a protein scaffold family concentrated in the PSD, are overexpression increases mushroom spine density (Haeckel et al.,
encoded by three genes, Shank1, Shank2 and Shank3, each giving 2008), cortactin overexpression causes elongation of spines
rise to multiple splice variants. Shank mutations have been linked (Hering and Sheng, 2003).
to autism and other neurodevelopmental/neuropsychiatric The above considerations lead us to suggest a model for
disorders (reviews: Grabrucker et al., 2011; Sala et al., 2015). activity-induced modification of spine morphology: under basal
All Shank isoforms have a similar molecular organization, with conditions, cytosolic Shanks within the spine are pegged to
specialized domains that can associate with GKAPs, Homers, the actin cytoskeleton through cortactin and held outside of
and other Shanks. Analysis of immunoEM data suggests that the pallium. During activity, CaMKII-mediated phosphorylation
Shanks at the PSD are organized into a proximal pool, close releases Shanks from cortactin. The Shank thus released
enough to the interface between the core and pallium to associate accumulates at the pallium, while cortactin exits the spine.
with GKAP, and a distal (deeper) pool that may be stabilized Shank accumulated at the pallium could then recruit Abp1
through association with Homers and/or with other Shanks. to promote remodeling of the actin cytoskeleton around the
Under excitatory conditions, Shanks accumulate preferentially in pallium.
the distal pool (Tao-Cheng et al., 2015).
Shanks promote maturation of dendritic spines and the ACTIN AND THE PSD
enlargement of spine heads (Sala et al., 2001), although the
precise mechanism remains unclear. Considering that changes There are several reports that filaments of F-actin, the primary
in spine shape and size involve reorganization of the actin cytoskeletal component of the spine, contact the PSD (Gulley and
cytoskeleton, it is likely that Shanks regulate spine morphology Reese, 1981; Landis and Reese, 1983; Fifková, 1985), especially
through interaction with actin. Indeed, Shanks associate with at its periphery (Burette et al., 2012), though these contacts are
three actin-regulating proteins, Insulin Receptor Substrate likely highly variable, considering the dynamic nature of F-actin.
Protein 53 (IRSp53), Abp1 and cortactin. While the molecular basis of the attachment of actin filaments
The actin binding protein IRSp53 (also called BAIAP2; see to the PSD remains uncertain, a number of PSD molecules are
review by Kang et al., 2016) is a major PSD component, with plausible candidates. As discussed above, two Shank binding
a molar ratio of IRSp53 to PSD-95 of 1:4 (Lowenthal et al., proteins, IRSp53 and Abp1, may provide a link between the PSD
2015). IRSp53 is located between layers containing Shank and and the actin cytoskeleton.
PSD-95, relatively close to the postsynaptic membrane (Burette Yet another point of interaction of actin with the PSD may
et al., 2014), where it may function as a linker between the actin be through the ubiquitous actin binding protein α-actinin. In
cytoskeleton and the plasma membrane (Scita et al., 2008; Burette vitro assays show that α-actinin can form a ternary complex
et al., 2014). Indeed, IRSp53 contains an actin-binding BAR-like with the PSD proteins, densin and CaMKII (Walikonis et al.,
domain that can induce changes in membrane curvature (Zhao 2001). Loss of α-actinin-2 in rat hippocampal neurons creates an
et al., 2011) and thus may be involved in activity-induced changes increased density of immature, filopodia-like protrusions that fail
in the curvature of the synapse (Burette et al., 2014). to mature into a mushroom-shaped spine during development
Abp1, which can associate simultaneously with actin and (Hodges et al., 2014).
Shanks, may link the actin cytoskeleton to the PSD (Qualmann
et al., 2004). Abp1 preferentially interacts with dynamic rather CONCLUSION
than static F-actin structures (Kessels et al., 2000) and more
Abp1 becomes incorporated into Shank-positive synapses during The pallium is an extension of the electron-dense PSD
activity (Qualmann et al., 2004). Overexpression of Abp1 core, demarcated by strong immunolabeling for two PSD
scaffold proteins, Shank and Homer. This specialized zone, of the actin cytoskeleton (Kim et al., 2015). Subsequent
sandwiched between the electron-dense PSD core and the accumulation of CaMKII at the pallium docks elements that
actin ‘‘spinoskeleton’’, is highly dynamic, exhibiting striking regulate protein turnover (Bingol et al., 2010; Thein et al.,
changes during synaptic activity. Under conditions of strong 2014).
excitation, the pallium becomes electron-dense, with the addition In summary, activation of CaMKII and its translocation to
of CaMKII and several other proteins (Table 1). Activation the pallium under excitatory conditions trigger a chain of events
and/or translocation of CaMKII is necessary for the recruitment poised to elicit profound effects on the organization of the PSD
of other components to the pallium. complex and of the actin cytoskeleton that could synchronize
The pallium can be viewed as a hub where several receptor trafficking with changes in spine morphology. We
proteins converge during activity. Accumulating evidence on speculate that the degree and duration of CaMKII accumulation
the movements of individual proteins suggests a mechanism for and activity at the pallium, promoted by different types of
concerted insertion of receptors to the PSD and re-organization excitatory stimuli, may determine the type and level of synaptic
of the actin spinoskeleton, both mediated by CaMKII. Upon modification.
synaptic stimulation, translocation and activation of CaMKII
cause SynGAP and AIDA-1 to move out of the PSD AUTHOR CONTRIBUTIONS
core and accumulate at the pallium (Yang et al., 2013;
Dosemeci et al., 2015, 2016). We propose that the movement All authors contributed to the writing of the manuscript.
of SynGAP (and perhaps also of AIDA-1) vacates ‘‘slots’’
on PSD-95, providing a window of opportunity for the ACKNOWLEDGMENTS
insertion of AMPA receptors. Simultaneous CaMKII-mediated
accumulation of Shanks at the pallium (Tao-Cheng et al., We would like to thank Dr. John Lisman for stimulating
2014b), on the other hand, may enable actin reorganization discussions and for a critical reading of the manuscript. This
around the PSD. CaMKII also acts as a dynamic structural work was supported by the Intramural Research Program of
element whose activity-induced translocation changes the the National Institute of Neurological Disorders and Stroke
molecular organization within the spine. Dissociation of (NINDS)/National Institutes of Health (NIH) and NINDS/NIH
CaMKII from F-actin causes destabilization and reorganization NS 039444 grant to RJW.
REFERENCES Chen, X., Winters, C., Azzam, R., Li, X., Galbraith, J. A., Leapman, R. D.,
et al. (2008). Organization of the core structure of the postsynaptic density.
Araki, Y., Zeng, M., Zhang, M., and Huganir, R. L. (2015). Rapid dispersion of Proc. Natl. Acad. Sci. U S A 105, 4453–4458. doi: 10.1073/pnas.08008
SynGAP from synaptic spines triggers AMPA receptor insertion and spine 97105
enlargement during LTP. Neuron 85, 173–189. doi: 10.1016/j.neuron.2014. Collins, M. O., Husi, H., Yu, L., Brandon, J. M., Anderson, C. N., Blackstock, W. P.,
12.023 et al. (2006). Molecular characterization and comparison of the components
Baron, M. K., Boeckers, T. M., Vaida, B., Faham, S., Gingery, M., Sawaya, M. R., and multiprotein complexes in the postsynaptic proteome. J. Neurochem. 97,
et al. (2006). An architectural framework that may lie at the core of the 16–23. doi: 10.1111/j.1471-4159.2005.03507.x
postsynaptic density. Science 311, 531–535. doi: 10.1126/science.1118995 Dakoji, S., Tomita, S., Karimzadegan, S., Nicoll, R. A., and Bredt, D. S.
Bingol, B., Wang, C. F., Arnott, D., Cheng, D., Peng, J., and Sheng, M. (2010). (2003). Interaction of transmembrane AMPA receptor regulatory proteins
Autophosphorylated CaMKIIα acts as a scaffold to recruit proteasomes to with multiple membrane associated guanylate kinases. Neuropharmacology 45,
dendritic spines. Cell 140, 567–578. doi: 10.1016/j.cell.2010.01.024 849–856. doi: 10.1016/s0028-3908(03)00267-3
Boeckers, T. M., Kreutz, M. R., Winter, C., Zuschratter, W., Smalla, K. H., DeRobertis, E. D. P. (1964). ‘‘International series of monographs on pure
Sanmarti-Vila, L., et al. (1999a). Proline-rich synapse-associated protein- and applied biology. Division, modern trends in physiological sciences,’’ in
1/cortactin binding protein 1 (ProSAP1/CortBP1) is a PDZ-domain protein Histophysiology of Synapses and Neurosecretion (Vol. 20), eds P. Alexander and
highly enriched in the postsynaptic density. J. Neurosci. 19, 6506–6518. Z. M. Bacq (Oxford: Pergamon Press), 34–42.
Boeckers, T. M., Winter, C., Smalla, K. H., Kreutz, M. R., Bockmann, J., Ding, J. D., Kennedy, M. B., and Weinberg, R. J. (2013). Subcellular organization
Seidenbecher, C., et al. (1999b). Proline-rich synapse-associated proteins of camkii in rat hippocampal pyramidal neurons. J. Comp. Neurol. 521,
ProSAP1 and ProSAP2 interact with synaptic proteins of the SAPAP/GKAP 3570–3583. doi: 10.1002/cne.23372
family. Biochem. Biophys. Res. Commun. 264, 247–252. doi: 10.1006/bbrc. Dosemeci, A., and Jaffe, H. (2010). Regulation of phosphorylation at the
1999.1489 postsynaptic density during different activity states of Ca2+ /calmodulin-
Burette, A. C., Lesperance, T., Crum, J., Martone, M., Volkmann, N., dependent protein kinase II. Biochem. Biophys. Res. Commun. 391, 78–84.
Ellisman, M. H., et al. (2012). Electron tomographic analysis of synaptic doi: 10.1016/j.bbrc.2009.10.167
ultrastructure. J. Comp. Neurol. 520, 2697–2711. doi: 10.1002/cne.23067 Dosemeci, A., Makusky, A. J., Jankowska-Stephens, E., Yang, X., Slotta, D. J.,
Burette, A. C., Park, H., and Weinberg, R. J. (2014). Postsynaptic distribution and Markey, S. P. (2007). Composition of the synaptic PSD-95 complex.
of IRSp53 in spiny excitatory and inhibitory neurons. J. Comp. Neurol. 522, Mol. Cell. Proteomics 6, 1749–1760. doi: 10.1074/mcp.M700040-
2164–2178. doi: 10.1002/cne.23526 MCP200
Chen, X., Nelson, C. D., Li, X., Winters, C. A., Azzam, R., Sousa, A. A., et al. (2011). Dosemeci, A., Tao-Cheng, J. H., Vinade, L., Winters, C. A., Pozzo-Miller, L.,
PSD-95 is required to sustain the molecular organization of the postsynaptic and Reese, T. S. (2001). Glutamate-induced transient modification of the
density. J. Neurosci. 31, 6329–6338. doi: 10.1523/JNEUROSCI.5968-10.2011 postsynaptic density. Proc. Natl. Acad. Sci. U S A 98, 10428–10432. doi: 10.
Chen, X., Vinade, L., Leapman, R. D., Petersen, J. D., Nakagawa, T., Phillips, T. M., 1073/pnas.181336998
et al. (2005). Mass of the postsynaptic density and enumeration of three key Dosemeci, A., Toy, D., Burch, A., Bayer, K. U., and Tao-Cheng, J. H. (2016).
molecules. Proc. Natl. Acad. Sci. U S A 102, 11551–11556. doi: 10.1073/pnas. CaMKII-mediated displacement of AIDA-1 out of the postsynaptic density
0505359102 core. FEBS Lett. doi: 10.1002/1873-3468.12334 [Epub ahead of print].
Dosemeci, A., Toy, D., Reese, T. S., and Tao-Cheng, J. H. (2015). AIDA-1 moves Martone, M. E., Jones, Y. Z., Young, S. J., Ellisman, M. H., Zivin, J. A., and Hu, B. R.
out of the postsynaptic density core under excitatory conditions. PLoS One (1999). Modification of postsynaptic densities after transient cerebral ischemia:
10:e0137216. doi: 10.1371/journal.pone.0137216 a quantitative and three-dimensional ultrastructural study. J. Neurosci. 19,
Dosemeci, A., Vinade, L., Winters, C. A., Reese, T. S., and Tao-Cheng, J. H. 1988–1997.
(2002). Inhibition of phosphatase activity prolongs NMDA-induced Naisbitt, S., Kim, E., Tu, J. C., Xiao, B., Sala, C., Valtschanoff, J., et al. (1999).
modification of the postsynaptic density. J. Neurocytol. 31, 605–612. doi: 10. Shank, a novel family of postsynaptic density proteins that binds to the NMDA
1023/A:1025735410738 receptor/PSD-95/GKAP complex and cortactin. Neuron 23, 569–582. doi: 10.
Erondu, N. E., and Kennedy, M. B. (1985). Regional distribution of type 1016/s0896-6273(00)80809-0
II Ca2+ /calmodulin-dependent protein kinase in rat brain. J. Neurosci. 5, Nathan, J. A., Kim, H. T., Ting, L., Gygi, S. P., and Goldberg, A. L. (2013).
3270–3277. Why do cellular proteins linked to K63-polyubiquitin chains not associate with
Fifková, E. (1985). Actin in the nervous system. Brain Res. 356, 187–215. doi: 10. proteasomes? EMBO J. 32, 552–565. doi: 10.1038/emboj.2012.354
1016/0165-0173(85)90012-8 Okamoto, K., Narayanan, R., Lee, S. H., Murata, K., and Hayashi, Y. (2007). The
Grabrucker, A. M., Schmeisser, M. J., Schoen, M., and Boeckers, T. M. (2011). role of CaMKII as an F-actin-bundling protein crucial for maintenance of
Postsynaptic ProSAP/Shank scaffolds in the cross-hair of synaptopathies. dendritic spine structure. Proc. Natl. Acad. Sci. U S A 104, 6418–6423. doi: 10.
Trends Cell Biol. 21, 594–603. doi: 10.1016/j.tcb.2011.07.003 1073/pnas.0701656104
Gulley, R. L., and Reese, T. S. (1981). Cytoskeletal organization at the postsynaptic Otmakhov, N., Tao-Cheng, J. H., Carpenter, S., Asrican, B., Dosemeci, A.,
complex. J. Cell Biol. 91, 298–302. doi: 10.1083/jcb.91.1.298 Reese, T. S., et al. (2004). Persistent accumulation of calcium/calmodulin-
Haeckel, A., Ahuja, R., Gundelfinger, E. D., Qualmann, B., and Kessels, M. M. dependent protein kinase II in dendritic spines after induction of NMDA
(2008). The actin-binding protein Abp1 controls dendritic spine morphology receptor-dependent chemical long-term potentiation. J. Neurosci. 24,
and is important for spine head and synapse formation. J. Neurosci. 28, 9324–9331. doi: 10.1523/JNEUROSCI.2350-04.2004
10031–10044. doi: 10.1523/jneurosci.0336-08.2008 Palay, S. L. (1956). Synapses in the central nervous system. J. Biophys. Biochem.
Hayashi, M. K., Tang, C., Verpelli, C., Narayanan, R., Stearns, M. H., Xu, R. M., Cytol. 2, 193–202. doi: 10.1083/jcb.2.4.193
et al. (2009). The postsynaptic density proteins Homer and Shank form a Petersen, J. D., Chen, X., Vinade, L., Dosemeci, A., Lisman, J. E., and Reese, T. S.
polymeric network structure. Cell 137, 159–171. doi: 10.1016/j.cell.2009.01.050 (2003). Distribution of postsynaptic density (PSD)-95 and Ca2+ /calmodulin-
Hell, J. W. (2014). CaMKII: claiming center stage in postsynaptic function and dependent protein kinase II at the PSD. J. Neurosci. 23, 11270–11278.
organization. Neuron 81, 249–265. doi: 10.1016/j.neuron.2013.12.024 Petralia, R. S., Sans, N., Wang, Y. X., and Wenthold, R. J. (2005). Ontogeny of
Hering, H., and Sheng, M. (2003). Activity-dependent redistribution and postsynaptic density proteins at glutamatergic synapses. Mol. Cell. Neurosci. 29,
essential role of cortactin in dendritic spine morphogenesis. J. Neurosci. 23, 436–452. doi: 10.1016/j.mcn.2005.03.013
11759–11769. Qualmann, B., Boeckers, T. M., Jeromin, M., Gundelfinger, E. D., and
Hodges, J. L., Vilchez, S. M., Asmussen, H., Whitmore, L. A., and Horwitz, A. R. Kessels, M. M. (2004). Linkage of the actin cytoskeleton to the postsynaptic
(2014). α-Actinin-2 mediates spine morphology and assembly of the post- density via direct interactions of Abp1 with the ProSAP/Shank family.
synaptic density in hippocampal neurons. PLoS One 9:e101770. doi: 10. J. Neurosci. 24, 2481–2495. doi: 10.1523/JNEUROSCI.5479-03.2004
1371/journal.pone.0101770 Racz, B., and Weinberg, R. J. (2004). The subcellular organization of cortactin in
Hu, B. R., Park, M., Martone, M. E., Fischer, W. H., Ellisman, M. H., and Zivin, J. A. hippocampus. J. Neurosci. 24, 10310–10317. doi: 10.1523/JNEUROSCI.2080-
(1998). Assembly of proteins to postsynaptic densities after transient cerebral 04.2004
ischemia. J. Neurosci. 18, 625–633. Rostaing, P., Real, E., Siksou, L., Lechaire, J. P., Boudier, T., Boeckers, T. M., et al.
Jacob, A. L., Jordan, B. A., and Weinberg, R. J. (2010). Organization of amyloid- (2006). Analysis of synaptic ultrastructure without fixative using high-pressure
beta protein precursor intracellular domain-associated protein-1 in the rat freezing and tomography. Eur. J. Neurosci. 24, 3463–3474. doi: 10.1111/j.1460-
brain. J. Comp. Neurol. 518, 3221–3236. doi: 10.1002/cne.22394 9568.2006.05234.x
Jordan, B. A., Fernholz, B. D., Khatri, L., and Ziff, E. B. (2007). Activity-dependent Rumbaugh, G., Adams, J. P., Kim, J. H., and Huganir, R. L. (2006). SynGAP
AIDA-1 nuclear signaling regulates nucleolar numbers and protein synthesis in regulates synaptic strength and mitogen-activated protein kinases in cultured
neurons. Nat. Neurosci. 10, 427–435. doi: 10.1038/nn1867 neurons. Proc. Natl. Acad. Sci. U S A 103, 4344–4351. doi: 10.1073/pnas.
Kang, J., Park, H., and Kim, E. (2016). IRSp53/BAIAP2 in dendritic 0600084103
spine development, NMDA receptor regulation and psychiatric disorders. Sala, C., Piëch, V., Wilson, N. R., Passafaro, M., Liu, G., and Sheng, M. (2001).
Neuropharmacology 100, 27–39. doi: 10.1016/j.neuropharm.2015.06.019 Regulation of dendritic spine morphology and synaptic function by Shank and
Kessels, M. M., Engqvist-Goldstein, A. E., and Drubin, D. G. (2000). Association Homer. Neuron 31, 115–130. doi: 10.1016/s0896-6273(01)00339-7
of mouse actin-binding protein 1 (mAbp1/SH3P7), an Src kinase target, Sala, C., Vicidomini, C., Bigi, I., Mossa, A., and Verpelli, C. (2015). Shank
with dynamic regions of the cortical actin cytoskeleton in response to Rac1 synaptic scaffold proteins: keys to understanding the pathogenesis of autism
activation. Mol. Biol. Cell 11, 393–412. doi: 10.1091/mbc.11.1.393 and other synaptic disorders. J. Neurochem. 135, 849–858. doi: 10.1111/jnc.
Kim, K., Lakhanpal, G., Lu, H. E., Khan, M., Suzuki, A., Hayashi, M. K., et al. 13232
(2015). A temporary gating of actin remodeling during synaptic plasticity Schnell, E., Sizemore, M., Karimzadegan, S., Chen, L., Bredt, D. S., and Nicoll, R. A.
consists of the interplay between the kinase and structural functions of CaMKII. (2002). Direct interactions between PSD-95 and stargazin control synaptic
Neuron 87, 813–826. doi: 10.1016/j.neuron.2015.07.023 AMPA receptor number. Proc. Natl. Acad. Sci. U S A 99, 13902–13907. doi: 10.
Kim, J. H., Liao, D., Lau, L. F., and Huganir, R. L. (1998). SynGAP: a synaptic 1073/pnas.172511199
RasGAP that associates with the PSD-95/SAP90 protein family. Neuron 20, Scita, G., Confalonieri, S., Lappalainen, P., and Suetsugu, S. (2008). IRSp53:
683–691. doi: 10.1016/s0896-6273(00)81008-9 crossing the road of membrane and actin dynamics in the formation of
Kim, E., Naisbitt, S., Hsueh, Y. P., Rao, A., Rothschild, A., Craig, A. M., et al. membrane protrusions. Trends Cell Biol. 18, 52–60. doi: 10.1016/j.tcb.2007.
(1997). GKAP, a novel synaptic protein that interacts with the guanylate kinase- 12.002
like domain of the PSD-95/SAP90 family of channel clustering molecules. J. Cell Shonesy, B. C., Jalan-Sakrikar, N., Cavener, V. S., and Colbran, R. J.
Biol. 136, 669–678. doi: 10.1083/jcb.136.3.669 (2014). CaMKII: a molecular substrate for synaptic plasticity and memory.
Landis, D. M., and Reese, T. S. (1983). Cytoplasmic organization in cerebellar Prog. Mol. Biol. Transl. Sci. 122, 61–87. doi: 10.1016/B978-0-12-420170-5.
dendritic spines. J. Cell Biol. 97, 1169–1178. doi: 10.1083/jcb.97.4.1169 00003-9
Lisman, J., Yasuda, R., and Raghavachari, S. (2012). Mechanisms of CaMKII Sugiyama, Y., Kawabata, I., Sobue, K., and Okabe, S. (2005). Determination
action in long-term potentiation. Nat. Rev. Neurosci. 13, 169–182. doi: 10. of absolute protein numbers in single synapses by a GFP-based calibration
1038/nrn3192 technique. Nat. Methods 2, 677–684. doi: 10.1038/nmeth783
Lowenthal, M. S., Markey, S. P., and Dosemeci, A. (2015). Quantitative mass Suzuki, T., Okumura-Noji, K., Tanaka, R., and Tada, T. (1994). Rapid
spectrometry measurements reveal stoichiometry of principal postsynaptic translocation of cytosolic Ca2+ /calmodulin-dependent protein kinase II into
density proteins. J. Proteome Res. 14, 2528–2538. doi: 10.1021/acs.jproteome. postsynaptic density after decapitation. J. Neurochem. 63, 1529–1537. doi: 10.
5b00109 1046/j.1471-4159.1994.63041529.x
Tao-Cheng, J. H., Dosemeci, A., Gallant, P. E., Smith, C., and Reese, T. (2010). of Ca2+ /calmodulin-dependent protein kinase II and α-actinin. J. Neurosci. 21,
Activity induced changes in the distribution of Shanks at hippocampal 423–433.
synapses. Neuroscience 168, 11–17. doi: 10.1016/j.neuroscience.2010. Xiao, B., Tu, J. C., Petralia, R. S., Yuan, J. P., Doan, A., Breder, C. D., et al. (1998).
03.041 Homer regulates the association of group 1 metabotropic glutamate receptors
Tao-Cheng, J. H., Thein, S., Yang, Y., Reese, T. S., and Gallant, P. E. (2014a). with multivalent complexes of homer-related, synaptic proteins. Neuron 21,
Homer is concentrated at the postsynaptic density and does not redistribute 707–716. doi: 10.1016/s0896-6273(00)80588-7
after acute synaptic stimulation. Neuroscience 266, 80–90. doi: 10.1016/j. Yang, Y., Tao-Cheng, J. H., Bayer, K. U., Reese, T. S., and Dosemeci, A.
neuroscience.2014.01.066 (2013). Camkii-mediated phosphorylation regulates distributions of Syngap-α1
Tao-Cheng, J. H., Yang, Y., Bayer, K. U., Reese, T. S., and Dosemeci, A. (2014b). and -α2 at the postsynaptic density. PLoS One 8:e71795. doi: 10.1371/journal.
NMDA-induced accumulation of Shank at the postsynaptic density is mediated pone.0071795
by CaMKII. Biochem. Biophys. Res. Commun. 450, 808–811. doi: 10.1016/j.bbrc. Yang, Y., Tao-Cheng, J. H., Reese, T. S., and Dosemeci, A. (2011). SynGAP moves
2014.06.049 out of the core of the postsynaptic density upon depolarization. Neuroscience
Tao-Cheng, J. H., Yang, Y., Reese, T. S., and Dosemeci, A. (2015). Differential 192, 132–139. doi: 10.1016/j.neuroscience.2011.06.061
distribution of Shank and GKAP at the postsynaptic density. PLoS One Zhao, H., Pykalainen, A., and Lappalainen, P. (2011). I-BAR domain proteins:
10:e0118750. doi: 10.1371/journal.pone.0118750 linking actin and plasma membrane dynamics. Curr. Opin. Cell Biol. 23, 14–21.
Thein, S., Tao-Cheng, J. H., Li, Y., Bayer, K. U., Reese, T. S., and Dosemeci, A. doi: 10.1016/j.ceb.2010.10.005
(2014). CaMKII mediates recruitment and activation of the deubiquitinase
CYLD at the postsynaptic density. PLoS One 9:e91312. doi: 10.1371/journal. Conflict of Interest Statement: The authors declare that the research was
pone.0091312 conducted in the absence of any commercial or financial relationships that could
Tu, J. C., Xiao, B., Naisbitt, S., Yuan, J. P., Petralia, R. S., Brakeman, P., et al. be construed as a potential conflict of interest.
(1999). Coupling of mGluR/Homer and PSD-95 complexes by the Shank family
of postsynaptic density proteins. Neuron 23, 583–592. doi: 10.1016/s0896- Copyright © 2016 Dosemeci, Weinberg, Reese and Tao-Cheng. This is an open-access
6273(00)80810-7 article distributed under the terms of the Creative Commons Attribution License
Valtschanoff, J. G., and Weinberg, R. J. (2001). Laminar organization of the (CC BY). The use, distribution and reproduction in other forums is permitted,
NMDA receptor complex within the postsynaptic density. J. Neurosci. 21, provided the original author(s) or licensor are credited and that the original
1211–1217. publication in this journal is cited, in accordance with accepted academic practice.
Walikonis, R. S., Oguni, A., Khorosheva, E. M., Jeng, C. J., Asuncion, F. J., and No use, distribution or reproduction is permitted which does not comply with these
Kennedy, M. B. (2001). Densin-180 forms a ternary complex with the α-subunit terms.
The age of genetically encoded voltage indicators (GEVIs) has matured to the point that
changes in membrane potential can now be observed optically in vivo. Improving the
signal size and speed of these voltage sensors has been the primary driving forces during
this maturation process. As a result, there is a wide range of probes using different
voltage detecting mechanisms and fluorescent reporters. As the use of these probes
transitions from optically reporting membrane potential in single, cultured cells to imaging
populations of cells in slice and/or in vivo, a new challenge emerges—optically resolving
the different types of neuronal activity. While improvements in speed and signal size are
still needed, optimizing the voltage range and the subcellular expression (i.e., soma only)
of the probe are becoming more important. In this review, we will examine the ability of
recently developed probes to report synaptic activity in slice and in vivo. The voltage-
sensing fluorescent protein (VSFP) family of voltage sensors, ArcLight, ASAP-1, and the
Edited by:
George Augustine, rhodopsin family of probes are all good at reporting changes in membrane potential, but
Nanyang Technological University, all have difficulty distinguishing subthreshold depolarizations from action potentials and
Singapore
detecting neuronal inhibition when imaging populations of cells. Finally, we will offer a few
Reviewed by:
Jason D. Shepherd, possible ways to improve the optical resolution of the various types of neuronal activities.
University of Utah, USA
Christian Wilms, Keywords: genetically-encoded voltage indicators, synaptic activity, optogenetics, brain slices, in vivo
Scientifica Ltd., UK
*Correspondence:
Bradley J. Baker
INTRODUCTION
[email protected]
†
As developers of genetically-encoded voltage indicators (GEVIs) we are often asked for our best
These authors have contributed
probe. Until recently, a good GEVI would have been any that gave a voltage-dependent, optical
equally to this work.
signal in mammalian cells (Dimitrov et al., 2007; Lundby et al., 2008, 2010; Perron et al., 2009a,b).
Received: 06 May 2016 Now the experimenter has several probes to choose from that differ in their voltage-dependencies,
Accepted: 25 July 2016 speed, signal size, and brightness (Akemann et al., 2012; Jin et al., 2012; Kralj et al., 2012; Han et al.,
Published: 05 August 2016 2013; St-Pierre et al., 2014; Zou et al., 2014; Gong et al., 2015; Piao et al., 2015; Abdelfattah et al.,
Citation: 2016). The combinations of these varying characteristics result in strengths and weaknesses of
Nakajima R, Jung A, Yoon B-J and every GEVI available. There is no perfect probe that can optically resolve action potentials, synaptic
Baker BJ (2016) Optogenetic activity, and neuronal inhibition in vivo. Some GEVIs will give large, voltage-dependent optical
Monitoring of Synaptic Activity with
signals but are very dim limiting their usefulness in vivo. Others will give large optical signals but
Genetically Encoded
Voltage Indicators. are very slow reducing their ability to resolve fast firing action potentials. So now, when asked
Front. Synaptic Neurosci. 8:22. which is the best probe, the answer is simply another question. What do you want to measure? To
doi: 10.3389/fnsyn.2016.00022 fit with the theme of this edition, we will assume that the answer to that question is synaptic activity.
enabling a nearly instantaneous response. The signal size was the probe activity (Flytzanis et al., 2014; St-Pierre et al.,
also large giving roughly a 70% ∆F/F optical signal per 100 mV 2014).
membrane depolarization (Kralj et al., 2012; Figure 1, Arch The weak fluorescence of Arch limits its use to single
(D95N)). cell in culture studies or to C. elegans (Kralj et al., 2012;
Arch excelled in speed and signal size but suffered from some Flytzanis et al., 2014) for two main reasons. The first is that
serious weaknesses. The first weakness was that the original the intrinsic fluorescence of higher order neuro-systems will
version had an associated, light-induced current. The D95N mask the fluorescence of Arch-type probes. The second is that
mutation drastically reduced this current but also resulted in a ∆F in addition to ∆F/F is an important characteristic of the
slower probe (Kralj et al., 2012). The second weakness was that GEVI when it comes to the signal to noise ratio. An example
it does not traffic well to the plasma membrane. Even with the of this is shown in Figure 2. The HEK cell in Figure 2 is
addition of endoplasmic reticulum and Golgi network release expressing a GEVI from which the ∆F and the ∆F/F traces
motifs, every image of a rhodopsin probe in the literature exhibits from three different light levels are shown (Lee et al., 2016).
high intracellular fluorescence (Kralj et al., 2012; Flytzanis et al., As can be seen from this comparison, a high ∆F/F value
2014; Gong et al., 2014, 2015; Hochbaum et al., 2014; Hou can be achieved by a large change in fluorescence or a small
et al., 2014). The third and most devastating weakness was change in fluorescence when the probe is dim. Notice the
that Arch is very dim. The best versions of Arch and related increased noise in trace 3, a telltale sign of poor expression/dim
probes are still at least 5× dimmer than the green fluorescent fluorescence.
protein (GFP) requiring exceptionally strong illumination, at An ingenious solution to compensate for the poor
least 700× the light intensity required for ASAP-1 to visualize fluorescence of the rhodopsin voltage probes was developed
FIGURE 2 | Varying light levels affect ∆F and ∆F/F values. (A) An HEK 293 cell expressing a single fluorescent protein (FP) based GEVI, Bongwoori, is shown in
resting light intensity (RLI), (B) the fluorescence traces of averaged ∆F (Fx -F0 ) values from three different regions with varying intensity, (C) the fluorescence traces
showing averaged ∆F/F values from the same regions in (B; modified with permission from Lee et al., 2016, Figure 6).
simultaneously by the Adam Cohen and Mark Schnitzer the brighter signal (Wilt et al., 2013). It is also difficult to only
laboratories. By fusing an FP to the rhodopsin protein, förster excite the donor chromophore and not the acceptor as well.
resonance energy transfer (FRET) enabled the rhodopsin These factors combined with the relatively low signal size of
chromophore to affect the fluorescence of the fused FP. This FRET-based probes prohibit any reliable absolute measurement
design reduced the excitation light intensity needed to visualize of membrane potential.
the GEVI while maintaining the speed of the optical response The second design involves a circularly-permuted fluorescent
since the voltage-sensing chromophore was still in the voltage protein (cpFP) attached to the VSD. Initial designs fused the
field. These probes could also cover different wavelengths since cpFP downstream of the VSD so that the chromophore was in
many different FPs could be fused to rhodopsin and give a the cytoplasm (Gautam et al., 2009; Barnett et al., 2012). Electrik
signal (Gong et al., 2014; Zou et al., 2014). While this made the PK gave very small signals less than 1% ∆F/F per 100 mV
rhodopsin probes better, the optical signal sometimes could only depolarization but were very fast having a tau under 2 ms. A
indirectly report neuronal activity by determining the frequency substantial increase in signal size was achieved when the cpFP
of the noise in the optical recording (See Supplementary Figure was placed between the S3 transmembrane segment and the S4
5 in Gong et al., 2014). Now, an exciting new version using the transmembrane segment of the VSD putting the chromophore
FP, mNeonGreen, has recently been reported (Gong et al., 2015). outside of the cell (St-Pierre et al., 2014). This probe, ASAP-1,
mNeonGreen is a very bright FP (Shaner et al., 2013) enabling is one of the better GEVIs giving a fast and robust optical signal
Ace2N-mNeon to resolve action potentials in vivo in both flies (tau = 1–2 ms and about 20% ∆F/F per 100 mV depolarization
and mice. in HEK cells). ASAP-1 has a very broad voltage range which
is virtually linear over much of the physiologically relevant
potentials of neurons.
Class II—VSD Containing GEVIs The third design of GEVIs that utilize a VSD simply fuses
The second class of GEVIs is also the oldest. The original GEVI, the FP at the carboxy-terminus which puts the chromophore in
Flash (Siegel and Isacoff, 1997), was the result of inserting the cytoplasm. During a systematic test of different FPs fused at
GFP downstream of the pore domain of the voltage-gated different linker lengths from the VSD done in collaboration by
potassium channel, Shaker. Like the rhodopsin-based probes, the Vincent Pieribone’s lab and Larry Cohen’s lab, a point mutation
first generation of VSD-based probes had significant drawbacks on the outside of the FP, Super Ecliptic pHlorin (Miesenbock
making them useless in mammalian cells (Baker et al., 2007). et al., 1998; Ng et al., 2002) converted an alanine to an aspartic
The main problem was that the GEVIs did not traffic to the acid improving the optical signal 15 fold from 1% ∆F/F to 15%
plasma membrane. In 2007, one of the biggest advancements in per 100 mV depolarization of the plasma membrane (Jin et al.,
GEVI development was achieved by the Knöpfel laboratory when 2012). This negative charge on the outside of the β-can seems to
they fused FPs to the VSD of the voltage-sensing phosphatase affect the fluorescence of a neighboring chromophore when S4
gene from Ciona intestinalis (Murata et al., 2005). This probe, moves since mutations that favor the monomeric form of the FP
voltage-sensing fluorescent protein (VSFP) 2.1 trafficked well reduce the voltage-dependent optical signal substantially (Kang
to the plasma membrane which resulted in the first voltage- and Baker, 2016). Further development of ArcLight has gotten
dependent optical signals from cultured neurons (Dimitrov et al., signals as high as 40% ∆F/F per 100 mV depolarization step (Han
2007). et al., 2013). While ArcLight has the drawback of being slow, its
Another issue with VSD-based GEVIs is that the brightness and signal size make it one of the better probes for
chromophore resides outside of the voltage field so the optical imaging in vivo and in slice. In 2015, two publications improving
signal relies on the conformational change of the VSD. These the speed of this sensor were published. One dramatically
probes are therefore generally slower than the rhodopsin-based improved the off rate called Arclightening but reduced the signal
probes, but a recently developed red-shifted GEVI is extremely size to under 10% ∆F/F per 100 mV depolarization (Treger
fast having taus under 1 ms (Abdelfattah et al., 2016). et al., 2015). The other, Bongwoori, improved the speed of
There are three different designs for GEVIs that utilize a the sensor and shifted the voltage response to more positive
VSD. The first design uses a FRET pair flanking the VSD. An potentials which improved the resolution of action potentials but
example is Butterfly 1.2 (Akemann et al., 2012). This probe decreased the signal size for synaptic potentials (Piao et al., 2015).
is somewhat slow and gives a very small optical signal, less The reduced optical signal response for sub-threshold potentials
than 3% ∆F/F per 100 mV depolarization. A butterfly style gives Bongwoori a better ‘‘contrast’’ for optically resolving action
probe that gives a faster and larger optical signal was developed potentials.
last year called Nabi (Sung et al., 2015). An advantage of A final design for researchers to consider when choosing a
FRET-based probes is that the ratiometric imaging can remove GEVI is the genetically encoded, hVOS (Chanda et al., 2005;
movement artifacts due to respiration and blood flow in vivo. Wang et al., 2010, 2012; Ghitani et al., 2015). First developed
Theoretically, a ratiometric measurement could also be used to in the Bezanilla lab, hVOS consists of an FP anchored to the
determine the absolute value of the membrane potential since plasma membrane with the addition of a small charged molecule,
the ratio is concentration independent. In practice, however, the DPA, that binds to the plasma membrane effectively acting as
relative fluorescence of the two chromophores differ substantially a fluorescent quencher. Since DPA is a lipophilic anion, the
resulting in a potential increase in the noise for the analysis of quenching agent will move from the outer surface of the plasma
the optical signal. Often the experimenter should only analyze membrane to the inner surface upon membrane depolarizations
FIGURE 4 | Continued
images of ArcLight demonstrate membrane localization (arrow). The FP,
mCherry, is localized to the nucleus to facilitate identification of transduced
neurons. (d–f) Low magnification of the olfactory bulb—onl, olfactory nerve
layer; gl, glomerular layer; epl, external plexiform layer; mcl, mitral cell layer. (g)
Wide-field resting fluorescence intensity. (h) Glomerular patterns of activation
after odor stimulation. (i) Odor-evoked optical signals from the region of
interest marked with a red circle in (g,h). (j) Six unfiltered single trials aligned to
the first sniff of odorant.
the pixels of the camera has changed. Under high magnification, is to use only the most abundant codons for rapid translation
a researcher can choose only pixels that correspond to regions of of the protein. This has been shown to be effective for
the cell that exhibited a fluorescent response. When imaging a cytoplasmic proteins. However, for membrane proteins slowing
population of cells, a pixel will be less likely to capture only the the translation to allow proper folding and insertion into the
responsive fluorescence. This situation is depicted in Figure 6. translocon may also be important (Norholm et al., 2012; Yu et al.,
When imaging a single cell, it is much easier to avoid the internal, 2015). Finally, limiting the expression of the GEVI to subcellular
non-responsive fluorescence and maximize the signal to noise components (i.e., the soma, dendrites, etc.) could also focus the
ratio. optical signal to the desired region of the neuron again improving
While the GEVIs currently available have shown significant the signal to noise ratio.
improvement in their ability to optically detect neuronal activity,
there is still much room for improvement. Refining the voltage- AUTHOR CONTRIBUTIONS
sensitivity will enable maximizing the optical signal. For instance,
a probe that only responded to hyperpolarization of the RN and AJ wrote the manuscript and contributed figures. B-JY
plasma membrane would make identifying the inhibited parts and BJB helped to write the manuscript.
of a neuronal circuit much easier. Improving the membrane
expression of the GEVI will decrease the nonresponsive ACKNOWLEDGMENTS
fluorescence in a population of cells, thereby improving the
signal to noise ratio. Most efforts to improve trafficking involve This work was funded by the Korea Institute of Science and
the addition of endoplasmic reticulum and Golgi release motifs. Technology (KIST) Institutional Program Multiscale Functional
Codon optimization is another approach which for membrane Connectomics, 2E26190 and KIST Institutional Program
proteins may be a misnomer. The idea of codon optimization 2E26170.
REFERENCES Emiliani, V., Cohen, A. E., Deisseroth, K., and Hausser, M. (2015). All-
optical interrogation of neural circuits. J. Neurosci. 35, 13917–13926. doi: 10.
Abdelfattah, A. S., Farhi, S. L., Zhao, Y., Brinks, D., Zou, P., Ruangkittisakul, A., 1523/jneurosci.2916-15.2015
et al. (2016). A bright and fast red fluorescent protein voltage indicator that Empson, R. M., Goulton, C., Scholtz, D., Gallero-Salas, Y., Zeng, H., and
reports neuronal activity in organotypic brain slices. J. Neurosci. 36, 2458–2472. Knöpfel, T. (2015). Validation of optical voltage reporting by the genetically
doi: 10.1523/jneurosci.3484-15.2016 encoded voltage indicator VSFP-Butterfly from cortical layer 2/3 pyramidal
Akemann, W., Mutoh, H., Perron, A., Park, Y. K., Iwamoto, Y., and neurons in mouse brain slices. Physiol. Rep. 3:e12468. doi: 10.14814/phy2.
Knöpfel, T. (2012). Imaging neural circuit dynamics with a voltage-sensitive 12468
fluorescent protein. J. Neurophysiol. 108, 2323–2337. doi: 10.1152/jn.00452. Flytzanis, N. C., Bedbrook, C. N., Chiu, H., Engqvist, M. K., Xiao, C., Chan,
2012 K. Y., et al. (2014). Archaerhodopsin variants with enhanced voltage-
Akemann, W., Sasaki, M., Mutoh, H., Imamura, T., Honkura, N., and sensitive fluorescence in mammalian and Caenorhabditis elegans neurons. Nat.
Knöpfel, T. (2013). Two-photon voltage imaging using a genetically Commun. 5:4894. doi: 10.1038/ncomms5894
encoded voltage indicator. Sci. Rep. 3:2231. doi: 10.1038/srep Gautam, S. G., Perron, A., Mutoh, H., and Knöpfel, T. (2009). Exploration
02231 of fluorescent protein voltage probes based on circularly permuted
Akemann, W., Song, C., Mutoh, H., and Knöpfel, T. (2015). Route to fluorescent proteins. Front. Neuroeng. 2:14. doi: 10.3389/neuro.16.
genetically targeted optical electrophysiology: development and applications of 014.2009
voltage-sensitive fluorescent proteins. Neurophotonics 2:021008. doi: 10.1117/1. Ghitani, N., Bayguinov, P. O., Ma, Y., and Jackson, M. B. (2015). Single-trial
nph.2.2.021008 imaging of spikes and synaptic potentials in single neurons in brain slices with
Antic, S. D., Empson, R. M., and Knöpfel, T. (2016). Voltage imaging to genetically encoded hybrid voltage sensor. J. Neurophysiol. 113, 1249–1259.
understand connections and functions of neuronal circuits. J. Neurophysiol. doi: 10.1152/jn.00691.2014
116, 135–152. doi: 10.1152/jn.00226.2016 Gong, Y., Huang, C., Li, J. Z., Grewe, B. F., Zhang, Y., Eismann, S., et al.
Baker, B. J., Lee, H., Pieribone, V. A., Cohen, L. B., Isacoff, E. Y., Knöpfel, T., et al. (2015). High-speed recording of neural spikes in awake mice and flies with
(2007). Three fluorescent protein voltage sensors exhibit low plasma membrane a fluorescent voltage sensor. Science 350, 1361–1366. doi: 10.1126/science.
expression in mammalian cells. J. Neurosci. Methods 161, 32–38. doi: 10.1016/j. aab0810
jneumeth.2006.10.005 Gong, Y., Wagner, M. J., Zhong Li, J., and Schnitzer, M. J. (2014). Imaging
Barnett, L., Platisa, J., Popovic, M., Pieribone, V. A., and Hughes, T. (2012). neural spiking in brain tissue using FRET-opsin protein voltage sensors. Nat.
A fluorescent, genetically-encoded voltage probe capable of resolving action Commun. 5:3674. doi: 10.1038/ncomms4674
potentials. PLoS One 7:e43454. doi: 10.1371/journal.pone.0043454 Han, Z., Jin, L., Platisa, J., Cohen, L. B., Baker, B. J., and Pieribone, V. A. (2013).
Cao, G., Platisa, J., Pieribone, V. A., Raccuglia, D., Kunst, M., and Nitabach, M. N. Fluorescent protein voltage probes derived from ArcLight that respond to
(2013). Genetically targeted optical electrophysiology in intact neural circuits. membrane voltage changes with fast kinetics. PLoS One 8:e81295. doi: 10.
Cell 154, 904–913. doi: 10.1016/j.cell.2013.07.027 1371/journal.pone.0081295
Carandini, M., Shimaoka, D., Rossi, L. F., Sato, T. K., Benucci, A., and Knöpfel, T. Hochbaum, D. R., Zhao, Y., Farhi, S. L., Klapoetke, N., Werley, C. A.,
(2015). Imaging the awake visual cortex with a genetically encoded voltage Kapoor, V., et al. (2014). All-optical electrophysiology in mammalian neurons
indicator. J. Neurosci. 35, 53–63. doi: 10.1523/jneurosci.0594-14.2015 using engineered microbial rhodopsins. Nat. Methods 11, 825–833. doi: 10.
Chanda, B., Blunck, R., Faria, L. C., Schweizer, F. E., Mody, I., and Bezanilla, F. 1038/nmeth.3000
(2005). A hybrid approach to measuring electrical activity in genetically Hou, J. H., Venkatachalam, V., and Cohen, A. E. (2014). Temporal dynamics of
specified neurons. Nat. Neurosci. 8, 1619–1626. doi: 10.1038/nn1558 microbial rhodopsin fluorescence reports absolute membrane voltage. Biophys.
Dimitrov, D., He, Y., Mutoh, H., Baker, B. J., Cohen, L., Akemann, W., et al. (2007). J. 106, 639–648. doi: 10.1016/j.bpj.2013.11.4493
Engineering and characterization of an enhanced fluorescent protein voltage Jin, L., Han, Z., Platisa, J., Wooltorton, J. R., Cohen, L. B., and Pieribone, V. A.
sensor. PLoS One 2:e440. doi: 10.1371/journal.pone.0000440 (2012). Single action potentials and subthreshold electrical events imaged in
neurons with a fluorescent protein voltage probe. Neuron 75, 779–785. doi: 10. Scott, G., Fagerholm, E. D., Mutoh, H., Leech, R., Sharp, D. J., Shew, W. L., et al.
1016/j.neuron.2012.06.040 (2014). Voltage imaging of waking mouse cortex reveals emergence of critical
Kang, B. E., and Baker, B. J. (2016). Pado, a fluorescent protein with proton channel neuronal dynamics. J. Neurosci. 34, 16611–16620. doi: 10.1523/jneurosci.3474-
activity can optically monitor membrane potential, intracellular pH and map 14.2014
gap junctions. Sci. Rep. 6:23865. doi: 10.1038/srep23865 Shaner, N. C., Lambert, G. G., Chammas, A., Ni, Y., Cranfill, P. J., Baird, M. A.,
Knöpfel, T. (2012). Genetically encoded optical indicators for the analysis et al. (2013). A bright monomeric green fluorescent protein derived from
of neuronal circuits. Nat. Rev. Neurosci. 13, 687–700. doi: 10.1038/ Branchiostoma lanceolatum. Nat. Methods 10, 407–409. doi: 10.1038/nmeth.
nrn3293 2413
Knöpfel, T., Gallero-Salas, Y., and Song, C. (2015). Genetically encoded voltage Siegel, M. S., and Isacoff, E. Y. (1997). A genetically encoded optical probe of
indicators for large scale cortical imaging come of age. Curr. Opin. Chem. Biol. membrane voltage. Neuron 19, 735–741. doi: 10.1016/s0896-6273(00)80955-1
27, 75–83. doi: 10.1016/j.cbpa.2015.06.006 Storace, D. A., Braubach, O. R., Jin, L., Cohen, L. B., and Sung, U. (2015a).
Kralj, J. M., Douglass, A. D., Hochbaum, D. R., Maclaurin, D., and Cohen, A. E. Monitoring brain activity with protein voltage and calcium sensors. Sci. Rep.
(2012). Optical recording of action potentials in mammalian neurons using a 5:10212. doi: 10.1038/srep10212
microbial rhodopsin. Nat. Methods 9, 90–95. doi: 10.1038/nmeth.1782 Storace, D., Rad, M. S., Han, Z., Jin, L., Cohen, L. B., Hughes, T., et al. (2015b).
Lee, S., Piao, H. H., Sepheri-Rad, M., Jung, A., Sung, U., Song, Y. K., et al. (2016). Genetically encoded protein sensors of membrane potential. Adv. Exp. Med.
Imaging membrane potential with two types of genetically encoded fluorescent Biol. 859, 493–509. doi: 10.1007/978-3-319-17641-3_20
voltage sensors. J. Vis. Exp. 4:e53566. doi: 10.3791/53566 Storace, D., Sepehri-Rad, M., Kang, B., Cohen, L. B., Hughes, T., and Baker, B. J.
Lundby, A., Akemann, W., and Knöpfel, T. (2010). Biophysical characterization (2016). Toward better genetically encoded sensors of membrane potential.
of the fluorescent protein voltage probe VSFP2.3 based on the voltage-sensing Trends Neurosci 39, 277–289. doi: 10.1016/j.tins.2016.02.005
domain of Ci-VSP. Eur. Biophys. J. 39, 1625–1635. doi: 10.1007/s00249-010- St-Pierre, F., Chavarha, M., and Lin, M. Z. (2015). Designs and sensing
0620-0 mechanisms of genetically encoded fluorescent voltage indicators. Curr. Opin.
Lundby, A., Mutoh, H., Dimitrov, D., Akemann, W., and Knöpfel, T. (2008). Chem. Biol. 27, 31–38. doi: 10.1016/j.cbpa.2015.05.003
Engineering of a genetically encodable fluorescent voltage sensor exploiting fast St-Pierre, F., Marshall, J. D., Yang, Y., Gong, Y., Schnitzer, M. J., and Lin, M. Z.
Ci-VSP voltage-sensing movements. PLoS One 3:e2514. doi: 10.1371/journal. (2014). High-fidelity optical reporting of neuronal electrical activity with
pone.0002514 an ultrafast fluorescent voltage sensor. Nat. Neurosci. 17, 884–889. doi: 10.
Maclaurin, D., Venkatachalam, V., Lee, H., and Cohen, A. E. (2013). Mechanism 1038/nn.3709
of voltage-sensitive fluorescence in a microbial rhodopsin. Proc. Natl. Acad. Sci. Sung, U., Sepehri-Rad, M., Piao, H. H., Jin, L., Hughes, T., Cohen, L. B., et al.
U S A. 110, 5939–5944. doi: 10.1073/pnas.1215595110 (2015). Developing fast fluorescent protein voltage sensors by optimizing FRET
Miesenbock, G., De Angelis, D. A., and Rothman, J. E. (1998). Visualizing interactions. PLoS One 10:e0141585. doi: 10.1371/journal.pone.0141585
secretion and synaptic transmission with pH-sensitive green fluorescent Ting, J. T., Daigle, T. L., Chen, Q., and Feng, G. (2014). Acute brain slice methods
proteins. Nature 394, 192–195. doi: 10.1038/28190 for adult and aging animals: application of targeted patch clamp analysis
Murata, Y., Iwasaki, H., Sasaki, M., Inaba, K., and Okamura, Y. (2005). and optogenetics. Methods Mol. Biol. 1183, 221–242. doi: 10.1007/978-1-4939-
Phosphoinositide phosphatase activity coupled to an intrinsic voltage sensor. 1096-0_14
Nature 435, 1239–1243. doi: 10.1038/nature03650 Treger, J. S., Priest, M. F., and Bezanilla, F. (2015). Single-molecule fluorimetry and
Mutoh, H., Akemann, W., and Knöpfel, T. (2012). Genetically engineered gating currents inspire an improved optical voltage indicator. Elife 4:e10482.
fluorescent voltage reporters. ACS. Chem. Neurosci. 3, 585–592. doi: 10. doi: 10.7554/elife.10482
1021/cn300041b Wachowiak, M., and Knöpfel, T. (2009). ‘‘Optical imaging of brain activity in vivo
Mutoh, H., and Knöpfel, T. (2013). Probing neuronal activities with genetically using genetically encoded probes,’’ in In Vivo Optical Imaging of Brain
encoded optical indicators: from a historical to a forward-looking perspective. Function, 2nd Edn., ed. R. D. Frostig (Boca Raton, FL: CRC Press/Taylor &
Pflugers. Arch. 465, 361–371. doi: 10.1007/s00424-012-1202-z Francis), 1–36.
Mutoh, H., Mishina, Y., Gallero-Salas, Y., and Knöpfel, T. (2015). Comparative Wang, D., Mcmahon, S., Zhang, Z., and Jackson, M. B. (2012). Hybrid voltage
performance of a genetically-encoded voltage indicator and a blue voltage sensor imaging of electrical activity from neurons in hippocampal slices
sensitive dye for large scale cortical voltage imaging. Front. Cell. Neurosci. 9:147. from transgenic mice. J. Neurophysiol. 108, 3147–3160. doi: 10.1152/jn.
doi: 10.3389/fncel.2015.00147 00722.2012
Ng, M., Roorda, R. D., Lima, S. Q., Zemelman, B. V., Morcillo, P., and Wang, D., Zhang, Z., Chanda, B., and Jackson, M. B. (2010). Improved probes for
Miesenbock, G. (2002). Transmission of olfactory information between three hybrid voltage sensor imaging. Biophys. J. 99, 2355–2365. doi: 10.1016/j.bpj.
populations of neurons in the antennal lobe of the fly. Neuron 36, 463–474. 2010.07.037
doi: 10.1016/s0896-6273(02)00975-3 Wilt, B. A., Fitzgerald, J. E., and Schnitzer, M. J. (2013). Photon shot noise limits
Norholm, M. H., Light, S., Virkki, M. T., Elofsson, A., Von Heijne, on optical detection of neuronal spikes and estimation of spike timing. Biophys.
G., and Daley, D. O. (2012). Manipulating the genetic code for J. 104, 51–62. doi: 10.1016/j.bpj.2012.07.058
membrane protein production: what have we learnt so far? Biochim. Yu, C. H., Dang, Y., Zhou, Z., Wu, C., Zhao, F., Sachs, M. S., et al. (2015).
Biophys. Acta. 1818, 1091–1096. doi: 10.1016/j.bbamem.2011. Codon usage influences the local rate of translation elongation to regulate co-
08.018 translational protein folding. Mol. Cell 59, 744–754. doi: 10.1016/j.molcel.2015.
Perron, A., Akemann, W., Mutoh, H., and Knöpfel, T. (2012). Genetically encoded 07.018
probes for optical imaging of brain electrical activity. Prog. Brain Res. 196, Zou, P., Zhao, Y., Douglass, A. D., Hochbaum, D. R., Brinks, D., Werley, C. A.,
63–77. doi: 10.1016/b978-0-444-59426-6.00004-5 et al. (2014). Bright and fast multicoloured voltage reporters via electrochromic
Perron, A., Mutoh, H., Akemann, W., Gautam, S. G., Dimitrov, D., Iwamoto, FRET. Nat. Commun. 5:4625. doi: 10.1038/ncomms5625
Y., et al. (2009a). Second and third generation voltage-sensitive fluorescent
proteins for monitoring membrane potential. Front. Mol. Neurosci. 2:5. doi: 10. Conflict of Interest Statement: The authors declare that the research was
3389/neuro.02.005.2009 conducted in the absence of any commercial or financial relationships that could
Perron, A., Mutoh, H., Launey, T., and Knöpfel, T. (2009b). Red-shifted voltage- be construed as a potential conflict of interest.
sensitive fluorescent proteins. Chem. Biol. 16, 1268–1277. doi: 10.1016/j.
chembiol.2009.11.014 Copyright © 2016 Nakajima, Jung, Yoon and Baker. This is an open-access article
Piao, H. H., Rajakumar, D., Kang, B. E., Kim, E. H., and Baker, B. J. (2015). distributed under the terms of the Creative Commons Attribution License (CC BY).
Combinatorial mutagenesis of the voltage-sensing domain enables the optical The use, distribution and reproduction in other forums is permitted, provided the
resolution of action potentials firing at 60 Hz by a genetically encoded original author(s) or licensor are credited and that the original publication in this
fluorescent sensor of membrane potential. J. Neurosci. 35, 372–385. doi: 10. journal is cited, in accordance with accepted academic practice. No use, distribution
1523/jneurosci.3008-14.2015 or reproduction is permitted which does not comply with these terms.
Department of Physiology and Program in Neuroscience, University of Maryland School of Medicine, Baltimore, MD, USA
FIGURE 4 | Nanoscale regional density of PSD-95 within the synapse correlates with probe diffusion coefficient. (A) Typical example of tracks and
PSD-95 positions. Tracks pseudo-colored purple, PSD-95 positions pseudo-colored by regional density. (B) Cumulative frequency distributions of probe Deff binned
in increasing regional densites of PSD-95 surrounding the subtracks (n = 448 subtracks in 0–5 PSD-95, 281 in 5–10, 159 in 10–15, 106 in 15–20, 59 in 20–25, 22 in
25–30, 24 in 30+). (C) PSD-95 regional density and the median Deff of the binding probe; relationship tested by linear regression.
FIGURE 5 | A non-binding transmembrane protein enters and slows within the synapse, but not as much as if it can bind PSD-95. (A) Typical examples
of tracked probes (purple) superimposed on shrPSD-95 positions (pseudo-colored on a scale of regional density same as Figure 4A; binding probe SEP-TM-Bind
(left), nonbinding probe SEP-TM-Nonbind (right). (B) Cumulative frequency distributions of the binding probe and nonbinding probe Deff within PSDs (n = 654
tracks/113 PSDs/11 cells/3 cultures for SEP-TM-Bind, 519/91/10/3 SEP-TM-Nonbind). (C) Cumulative frequency distributions of the binding probe and nonbinding
probe Deff outside of PSDs (n = 3349 tracks for SEP-TM-Bind, 4470 SEP-TM-Nonbind). (D) (Left) PSD area and median Deff within each of the PSDs; linear
regression test (n = 91 PSDs/10 cells/3 cultures of SEP-TM-Nonbind, same as in Figure 3 for SEP-TM-Bind). (Right) Overall synaptic PSD-95 density and median
Deff within each of the PSDs; linear regression test (n = as in left panel).
probe-scaffold binding, we performed PALM-PAINT on a probe binding probe (Figure 5B). On the other hand, the extrasynaptic
variant that cannot bind to PSD-95 (SEP-TM-Nonbind from diffusion of the nonbinding probe was not different from that of
Li et al., 2016). Interestingly, SEP-TM-Nonbind still entered the binding probe (Figure 5C). It is interesting to note that using
synapses and diffused within them, but did not enrich within sptPALM (i.e., by photoactivating and tracking an mEos3 fusion
the PSD nearly as greatly as SEP-TM-Bind (Figure 5A and see protein rather than using an anti-GFP PAINT approach as here)
also Li et al., 2016). Notably, the diffusion of this nonbinding there was a very similar mobility differential between the binding
probe within the synapse was dramatically faster than that of the and nonbinding probes despite a substantial absolute difference
in Deff arising from the poorer localization precision of the fusion that despite the lack of a PSD-95-binding motif, the probe still
protein compared to the organic dye (Li et al., 2016). Thus, diffused more slowly within higher density regions of the PSD
the mobility difference between the two probes is quite robust. (Figure 6A). The effect appeared to saturate at low Deff since the
Intriguingly, the shoulder-like shape of the cumulative Deff mobility of these slowly moving molecules is below our detection
distribution suggests that probes undergo multiple influences limit (0.003 µm2 /s).
on their diffusion within the synapse. However, neither PSD If steric obstruction influences probe mobility, it may
area nor whole-synapse PSD-95 density correlated with probe influence the overall pattern of probe position within the synapse.
diffusion within the PSD (Figure 5D). We thus compared the fraction of subtracks found in different
By labeling SEP-tagged TM proteins with nanobodies, we add regional PSD-95 densities (Figure 6B). Interestingly, though the
minimal but appreciable bulk to the extracellular domain of the binding and the nonbinding probes were distributed similarly
diffusing entity. SEP is fused to the extracellular domain of the through most density values, the nonbinding probes appeared
TM probes, adding approximately 3–5 nm of bulk; nanobody to be preferentially excluded from the highly dense subregions
labeling of the GFP adds an additional 3–5 nm of bulk. While (i.e., >25 regional molecules) of the synapse. Because of the
additional extracellular bulk has previously been shown to slow small number of subtracks in these bins, however, this difference
receptor diffusion (Groc et al., 2007), we expect the effect to be was not significant. Note also that the distribution of tracked
identical for both the binding and the nonbinding probes and molecules may not completely faithfully represent the total
thus not change our conclusions about the effect of postsynaptic steady-state distribution of probe molecules, since molecules
steric hindrance. immobilized in the synapse for long periods are less likely to be
We next considered whether the regional density of PSD-95 recognized by a nanobody and be tracked by uPAINT.
immediately surrounding the nonbinding probe can sterically It appeared that regional densities of PSD-95 higher than
control the probe diffusion. To test this, we first subdivided the 10 had minimal effect on diffusion of the nonbinding probe
tracks into subtracks and binned them into increasing regional in the Deff range below our detection limit (Figure 6C),
densities of PSD-95 molecules, as in Figure 4B. This revealed whereas the regional density of PSD-95 linearly correlated with
FIGURE 6 | Subsynaptic regional density of PSD-95 influences the mobility of a probe that does not bind PSD-95. (A) Cumulative frequency distributions
of the nonbinding probe Deff binned in increasing regional densites of PSD-95 surrounding the subtracks (n = 480 subtracks in 0–5 PSD-95, 204 in 5–10, 148 in
10–15, 81 in 15–20, 32 in 20–25, 12 in 25–30, 13 in 30+). (B) Fraction of subtracks in different regional densities of PSD-95. (C) PSD-95 regional density and the
median Deff of the nonbinding probe. (D) Fraction of tracks with subsegments that were slowed below the detection limit per PSD (n = 113 PSDs for SEP-TM-Bind,
91 SEP-TM-Nonbind; ∗ p = 0.0123 Mann-Whitney U test). (E) (Left) Cartoon highlighting the PSD-95 molecules surrounding subtrack durations that diffused below
the detection limit. The open and closed purple circles indicate the beginning and end of a track, the circles pseudo-colored by regional density were within 30 nm of
sub-detection limit subtracks. We calculated the median regional density of PSD-95 of all subtrack durations that diffused below the detection limit with every PSD.
(Right) Median regional density of PSD-95 surrounding subtracks that were below the detection limit (n same as in D; ∗ p = 0.0325 K-S test). (F) Median regional
density of PSD-95 surrounding subtracks that above the detection limit (n same as in D; ns, Not significant, p = 0.158 K-S test).
receptors within the synapse; and macromolecular crowding in with sites of neurotransmitter release, as indicated by their
combination with binding stabilizes the receptors at subsynaptic transsynaptic alignment with RIM1/2 molecules which in
domains highly packed with other proteins important for turn correlate with presynaptic vesicle fusion locations (Tang
synaptic transmission. et al., in press). Thus, we speculate that the enhanced
Interestingly, though the TM binding probe is much smaller crowding within these high-density subdomains will slow
than AMPARs, its distribution of Deff within the synapse and help limit the escape of receptors from points in the
appeared as slow as, or even slower than, AMPAR diffusion synapse where they are most likely to be activated during
measured in other studies using dye-conjugated antibodies or neurotransmission.
quantum-dot-conjugated nanobodies (Nair et al., 2013; Cai et al., Crowding within high-density subdomains is likely not
2014). It is possible that these various labeling approaches due only to postsynaptic scaffolding proteins. Indeed,
preferentially sample receptors that exit the synapse and diffuse transmembrane adhesion molecules, which associate with
more freely in the perisynaptic space, which may tend to one another across the synaptic cleft, will enhance crowding
obscure real differences in the relative numbers of low-Deff further in the extracellular, transmembrane, and intracellular
molecules that occupy the synapse for long periods. On the domains. The distribution of these adhesion molecules may
other hand, taken at face value, the similarity is consistent regulate the alignment between neurotransmitter receptors and
with our previous report using sptPALM that a substantial release sites, though this is not known. The TM probes employed
and nearly identical fraction of mEos-tagged AMPARs and here share commonalities with some of the intercellular adhesion
binding probes diffused with Deff < 0.02 µm2 /s (Li et al., molecules in size (e.g., single-pass TM proteins) or the ability
2016), further supporting the notion that much of the synapse to bind PSD-95, yet synaptic adhesion proteins display quite
is so crowded it stabilizes and organizes both large and small divergent patterns of expression within the synapse. SynCAM
TM proteins. is distributed in clusters surrounding the border of PSD-95
Previously, we demonstrated using partial synapse molecules (Perez de Arce et al., 2015), as are some members
Fluorescence Recovery after Photobleaching (FRAP), of the cadherin/catenin system (Uchida et al., 1996). On the
smtPALM alone, and uPAINT alone that the protein bulk other hand, neuroligin1 and LRRTM2 are inside the synapse and
of the TM probe decreased its diffusion within the synapse. distributed in clusters reminiscent of PSD-95 (Chamma et al.,
However, previous approaches were limited in ability to 2016).
assess how the complex structure within the PSD could Surprisingly, LRRTM2, though smaller in both extracellular
influence the mobility and organization of TM proteins in and intracellular length than neuroligin1, in fact diffuses slower
a living synapse. Results from PALM-PAINT indicate that a than neuroligin1 (Chamma et al., 2016). Moreover, LRRTM2
particular degree of postsynaptic crowding, which we estimate is more compactly distributed and in the synaptic center than
as 5000 molecules/µm2 , can be sufficient to stabilize TM the larger neuroligin1, suggesting that LRRTM2 is more likely
protein diffusion in the absence of binding. This density associated with dense regions of PSD-95 which are often found
translates to an average ∼14 nm inter-PSD-95 distance, very near the center of the PSD (MacGillavry et al., 2013; Tang
similar to the mean nearest-neighbor distance of ∼13 nm et al., in press). This paradoxical result confirms that the
between the ‘‘vertical filaments’’ corresponding to PSD-95 arrangements of TM proteins cannot be predicted by their
as measured in EM tomography (Chen et al., 2008). Note bulk alone. Interestingly, however, AMPARs associate with
that we deduce this as an average spacing, but could not LRRTM2 through their extracellular domains (de Wit et al.,
directly measure it around individual moving probes. The 2009), potentially regulating the mobility of each protein by
similarity between these values suggests that rather subtle the addition of further bulk and protein-protein interactions.
variations in scaffold density across the lateral extent of the Further experiments are needed to tease out how the various
synapse could change TM protein mobility substantially. This interactions on diverse synaptic TM protein species dictate one
high density packing is similar to what has been measured another’s spatial arrangement.
for AMPA receptors (e.g., 2000–4000/µm2 , see Levet et al., The diffusion of TM proteins appeared complicated outside
2015). Indeed, receptor-scaffold binding may facilitate the of the synapse particularly within few hundred nanometers from
assembly of this tight packing. Though the fractional time the border of the PSD. In some cases, the nonbinding probe
synaptic AMPARs spend bound to PSD-95 is not known, moves over a large area in the extrasynaptic regions, but in other
macromolecular crowding is likely to augment maintenance cases, they can appear confined or immobilized as the binding
of this architecture once assembled, because receptors in probes. Proteins other than scaffolding proteins can certainly
crowded areas that dissociate from scaffolds will face a longer affect the mobility of these probes outside the synapse. Some
escape time from the region and thus are more likely to factors could affect both probes roughly equally: regions with
rebind PSD-95. high density of endocytic adaptor molecules (Blanpied et al.,
If domains of high PSD-95 density tend to accumulate 2002; Petralia et al., 2003; Racz et al., 2004), zones of dense
not only probes such as used here but also receptors, cortical cytoskeleton (He et al., 2016), sites of plasma membrane-
then their impact on synaptic transmission will depend on ER apposition (Spacek and Harris, 1997). In addition, puncta
where they are with respect to sites of neurotransmitter adherens or clusters of adhesion molecules (Perez de Arce et al.,
release (MacGillavry et al., 2011). Recently, we have found 2015) may slow transit of all TM proteins. Less frequently,
that nanoclusters of PSD-95 frequently align transsynaptically undetected regions could exist that would selectively affect the
binding probe. For example, small and low-density regions below alignment of the uPAINT and PALM data was subject to residual
the PSD detection criteria (<10 molecules and <30 nm in errors that may have diminished a larger underlying effect. The
diameter) could be situated within nanometers of the detected two color channels faced an alignment error of ∼6 nm, which
PSD border. These could come from small segments of multi- would somewhat blur our measurement of regional PSD-95
segmented PSDs (Spacek and Hartmann, 1983; Stewart et al., density around individual tracked locations. In addition, the
2005), but these types of PSDs are rare. In addition, though uPAINT data is subject to error stemming from the finite
the density of PSD-95 outside the PSD border is quite low precision of individual localizations. The Atto647N we used
(Zhang and Diamond, 2009; Perez de Arce et al., 2015), its for tracking is a relatively bright organic dye and helps to
extrasynaptic mobility has not been measured and there may maximize this precision and thus minimize error in the estimate
be enough PSD-95 in the perisynaptic region to bind and of Deff . However, brighter, longer-lasting fluorophores could
immobilize the binding probes. We speculate that the overall be advantageous. Nanobody-labeled small quantum dots (Wang
extrasynaptic diffusion of both probes are not different because et al., 2014) have been used to track AMPARs in and around
few extrasynaptic regions selectively affect the binding probe but synapses, and have the additional advantage of being so bright
not the nonbinding probe. as to facilitate tracking in 3D (Cai et al., 2014). However, 3D
mapping of the PSD would require high localization numbers
and longer imaging durations (Legant et al., 2016; and see below),
Role of Crowding in Synaptic Plasticity
and the z resolution normally obtainable without 4pi detection is
The many types and time scales of ongoing and triggered
usually worse than 100 nm for fluorescent proteins, making this
PSD plasticity that have been documented (Okabe et al., 1999;
difficult to implement.
Blanpied et al., 2008; MacGillavry et al., 2013; Bosch et al.,
In our application of PALM-PAINT, there was only limited
2014) suggest that PSD reorganization during plasticity will
temporal relationship between individual tracks (generally
affect accumulation not just of TM proteins like AMPARs
lasting <1 s and the PALM map (aggregated over the imaging
but additional molecules contributing to crowding as well. For
session of generally 4–6 min)). Though lateral drift was corrected
instance, PSD-95 content in spines has been shown to decrease
during this time (to an error we estimated as <10 nm), ongoing
transiently after an LTP induction protocol in hippocampal slice
morphing and internal reorganization of the PSD (Kerr and
cultures (Steiner et al., 2008). This may directly lead to loss of
Blanpied, 2012; MacGillavry et al., 2013) presumably degraded
AMPARs as their binding partners are lost. However, the loss
many details of the PSD-95 distribution in our final images.
of PSD-95 may decrease crowding, which could prompt a net
The reduced precision in capturing the true regional density
loss of even TM proteins with minimal direct binding to PSD-95
of PSD-95 molecules would diminish the difference we saw
(e.g., desensitized AMPARs uncoupled from stargazin Constals
between probes in different regional densities, and also reduce
et al., 2015). This further loss of molecular crowders could then
the difference between binding and nonbinding probe. Further,
further facilitate receptor exit. However, whether the transient
probes in a similar subsynaptic space but tracked early vs. late
loss of spine PSD-95 reflects a disruption of high-density areas
in the mapping might have not truly experienced the same
of the synapse is unknown.
degree of steric hindrance. However, the differences we observed
On the other hand, speculating further, if the transient
were robust even in the face of these errors. Ideally, to capture
decrease in synaptic PSD-95 during LTP induction reflects
the true effect size, one would need to monitor lateral drift
primarily loss from the PSD edge, and thus is not correlated
continuously (Bon et al., 2015) and achieve more rapid mapping
with significant de-crowding at high-density regions, then
(Huang et al., 2013). However, in structures with low protein
the continued presence of additional nonbinding proteins at
copy number, a large fraction of the proteins must be mapped
these high-density regions could in fact obstruct the entry of
to achieve statistical reliability (MacGillavry et al., 2013; Legant
AMPARs to these regions and limit the changes in AMPARs
et al., 2016), which may preclude time-lapse imaging except if the
level, at least over certain kinetic phases. Thus, it would be
protein exchange rate is high compared to the photobleaching
tempting to speculate in this case that the nonbinding proteins
rate induced by imaging.
could affect LTP induction kinetics but not maintenance, as
We hope the combined approach of PALM-PAINT will
morphing dynamics and internal mixing of the PSD would
help answer many key questions regarding synapse architecture
eventually enable synapses to reach their new steady state
and plasticity. One key issue is what mechanisms assemble
capacity of AMPARs on a time scale of minutes. Resolving these
the particular organization of PSD-95, a pattern that appears
many possibilities in the future will require close examination
to dictate receptor number and position (Opazo et al., 2012).
of the kinetics of protein redistribution and exchange during
One possibility is that the more deeply positioned multi-domain
plasticity.
proteins in the PSD, such as the Shank and GKAP families
(Valtschanoff and Weinberg, 2001; Dani et al., 2010), may
Advantages and Disadvantages establish a platform of loose spacing with which the more
of PALM-PAINT superficial proteins such as PSD-95 may interact (Chen et al.,
PALM of the PSD border improves discrimination of those 2008). Interestingly, in this case, a close interaction of the
molecules definitively within the synapse proper. However, deeper PSD with cytoskeleton (Frost et al., 2010; MacGillavry
we suspect that the effect of crowding may have been et al., 2016) may thus provide a link between activity-dependent
underestimated in our analysis because spatial and temporal plasticity of spine and PSD structure. Alternatively, cleft-resident
REFERENCES de Wit, J., Sylwestrak, E., O’Sullivan, M. L., Otto, S., Tiglio, K., Savas, J. N., et al.
(2009). LRRTM2 interacts with Neurexin1 and regulates excitatory synapse
Annibale, P., Vanni, S., Scarselli, M., Rothlisberger, U., and Radenovic, A. (2011). formation. Neuron 64, 799–806. doi: 10.1016/j.neuron.2009.12.019
Identification of clustering artifacts in photoactivated localization microscopy. Dupuis, J. P., Ladépêche, L., Seth, H., Bard, L., Varela, J., Mikasova, L., et al.
Nat. Methods 8, 527–528. doi: 10.1038/nmeth.1627 (2014). Surface dynamics of GluN2B-NMDA receptors controls plasticity of
Bats, C., Groc, L., and Choquet, D. (2007). The interaction between Stargazin maturing glutamate synapses. EMBO J. 33, 842–861. doi: 10.1002/embj.2013
and PSD-95 regulates AMPA receptor surface trafficking. Neuron 53, 719–734. 86356
doi: 10.1016/j.neuron.2007.01.030 Ehlers, M. D., Heine, M., Groc, L., Lee, M. C., and Choquet, D. (2007). Diffusional
Betzig, E., Patterson, G. H., Sougrat, R., Lindwasser, O. W., Olenych, S., trapping of GluR1 AMPA receptors by input-specific synaptic activity. Neuron
Bonifacino, J. S., et al. (2006). Imaging intracellular fluorescent proteins 54, 447–460. doi: 10.1016/j.neuron.2007.04.010
at nanometer resolution. Science 313, 1642–1645. doi: 10.1126/science.11 Erlenhardt, N., Yu, H., Abiraman, K., Yamasaki, T., Wadiche, J. I., Tomita, S.,
27344 et al. (2016). Porcupine controls hippocampal AMPAR levels, composition and
Blanpied, T. A., Kerr, J. M., and Ehlers, M. D. (2008). Structural plasticity with synaptic transmission. Cell Rep. 14, 782–794. doi: 10.1016/j.celrep.2015.12.078
preserved topology in the postsynaptic protein network. Proc. Natl. Acad. Sci. Freche, D., Pannasch, U., Rouach, N., and Holcman, D. (2011). Synapse geometry
U S A 105, 12587–12592. doi: 10.1073/pnas.0711669105 and receptor dynamics modulate synaptic strength. PLoS One 6:e25122. doi: 10.
Blanpied, T. A., Scott, D. B., and Ehlers, M. D. (2002). Dynamics and regulation 1371/journal.pone.0025122
of clathrin coats at specialized endocytic zones of dendrites and spines. Neuron Frick, M., Schmidt, K., and Nichols, B. J. (2007). Modulation of lateral diffusion
36, 435–449. doi: 10.1016/s0896-6273(02)00979-0 in the plasma membrane by protein density. Curr. Biol. 17, 462–467. doi: 10.
Bon, P., Bourg, N., Lécart, S., Monneret, S., Fort, E., Wenger, J., et al. 1016/j.cub.2007.01.069
(2015). Three-dimensional nanometre localization of nanoparticles to enhance Frost, N. A., Shroff, H., Kong, H., Betzig, E., and Blanpied, T. A. (2010). Single-
super-resolution microscopy. Nat. Commun. 6:7764. doi: 10.1038/ncomm molecule discrimination of discrete perisynaptic and distributed sites of actin
s8764 filament assembly within dendritic spines. Neuron 67, 86–99. doi: 10.1016/j.
Bosch, M., Castro, J., Saneyoshi, T., Matsuno, H., Sur, M., and Hayashi, Y. (2014). neuron.2010.05.026
Structural and molecular remodeling of dendritic spine substructures during Fukata, Y., Dimitrov, A., Boncompain, G., Vielemeyer, O., Perez, F., and
long-term potentiation. Neuron 82, 444–459. doi: 10.1016/j.neuron.2014. Fukata, M. (2013). Local palmitoylation cycles define activity-regulated
03.021 postsynaptic subdomains. J. Cell Biol. 202, 145–161. doi: 10.1083/jcb.2013
Broadhead, M. J., Horrocks, M. H., Zhu, F., Muresan, L., Benavides-Piccione, R., 02071
DeFelipe, J., et al. (2016). PSD95 nanoclusters are postsynaptic building blocks Giannone, G., Hosy, E., Levet, F., Constals, A., Schulze, K., Sobolevsky, A. I., et al.
in hippocampus circuits. Sci. Rep. 6:24626. doi: 10.1038/srep24626 (2010). Dynamic superresolution imaging of endogenous proteins on living
Cai, E., Ge, P., Lee, S. H., Jeyifous, O., Wang, Y., Liu, Y., et al. (2014). Stable small cells at ultra-high density. Biophys. J. 99, 1303–1310. doi: 10.1016/j.bpj.2010.
quantum dots for synaptic receptor tracking on live neurons. Angew. Chem. 06.005
Int. Ed Engl. 53, 12484–12488. doi: 10.1002/anie.201405735 Groc, L., Lafourcade, M., Heine, M., Renner, M., Racine, V., Sibarita, J. B., et al.
Chamma, I., Letellier, M., Butler, C., Tessier, B., Lim, K. H., Gauthereau, I., (2007). Surface trafficking of neurotransmitter receptor: comparison between
et al. (2016). Mapping the dynamics and nanoscale organization of synaptic single-molecule/quantum dot strategies. J. Neurosci. 27, 12433–12437. doi: 10.
adhesion proteins using monomeric streptavidin. Nat. Commun. 7:10773. 1523/jneurosci.3349-07.2007
doi: 10.1038/ncomms10773 Harris, K. M., and Stevens, J. K. (1989). Dendritic spines of CA 1 pyramidal
Chen, X., Winters, C., Azzam, R., Li, X., Galbraith, J. A., Leapman, R. D., et al. cells in the rat hippocampus: serial electron microscopy with reference to their
(2008). Organization of the core structure of the postsynaptic density. Proc. biophysical characteristics. J. Neurosci. 9, 2982–2997.
Natl. Acad. Sci. U S A 105, 4453–4458. doi: 10.1073/pnas.0800897105 Harris, K. M., and Weinberg, R. J. (2012). Ultrastructure of synapses in the
Cho, C. H., St-Gelais, F., Zhang, W., Tomita, S., and Howe, J. R. (2007). Two mammalian brain. Cold Spring Harb. Perspect. Biol. 4:a005587. doi: 10.
families of TARP isoforms that have distinct effects on the kinetic properties 1101/cshperspect.a005587
of AMPA receptors and synaptic currents. Neuron 55, 890–904. doi: 10.1016/j. He, J., Zhou, R., Wu, Z., Carrasco, M. A., Kurshan, P. T., Farley, J. E., et al. (2016).
neuron.2007.08.024 Prevalent presence of periodic actin-spectrin-based membrane skeleton in a
Choquet, D., and Triller, A. (2013). The dynamic synapse. Neuron 80, 691–703. broad range of neuronal cell types and animal species. Proc. Natl. Acad. Sci.
doi: 10.1016/j.neuron.2013.10.013 U S A 113, 6029–6034. doi: 10.1073/pnas.1605707113
Constals, A., Penn, A. C., Compans, B., Toulmé, E., Phillipat, A., Marais, S., Hess, S. T., Girirajan, T. P., and Mason, M. D. (2006). Ultra-high resolution
et al. (2015). Glutamate-induced AMPA receptor desensitization increases imaging by fluorescence photoactivation localization microscopy. Biophys. J.
their mobility and modulates short-term plasticity through unbinding from 91, 4258–4272. doi: 10.1529/biophysj.106.091116
Stargazin. Neuron 85, 787–803. doi: 10.1016/j.neuron.2015.01.012 Hoze, N., Nair, D., Hosy, E., Sieben, C., Manley, S., Herrmann, A., et al.
Dani, A., Huang, B., Bergan, J., Dulac, C., and Zhuang, X. (2010). Superresolution (2012). Heterogeneity of AMPA receptor trafficking and molecular interactions
imaging of chemical synapses in the brain. Neuron 68, 843–856. doi: 10.1016/j. revealed by superresolution analysis of live cell imaging. Proc. Natl. Acad. Sci.
neuron.2010.11.021 U S A 109, 17052–17057. doi: 10.1073/pnas.1204589109
Huang, F., Hartwich, T. M., Rivera-Molina, F. E., Lin, Y., Duim, W. C., adhesive nanodomains. Neuron 88, 1165–1172. doi: 10.1016/j.neuron.2015.
Long, J. J., et al. (2013). Video-rate nanoscopy using sCMOS camera-specific 11.011
single-molecule localization algorithms. Nat. Methods 10, 653–658. doi: 10. Petralia, R. S., Wang, Y. X., and Wenthold, R. J. (2003). Internalization at
1038/nmeth.2488 glutamatergic synapses during development. Eur. J. Neurosci. 18, 3207–3217.
Huganir, R. L., and Nicoll, R. A. (2013). AMPARs and synaptic plasticity: doi: 10.1111/j.1460-9568.2003.03074.x
the last 25 years. Neuron 80, 704–717. doi: 10.1016/j.neuron.2013. Racz, B., Blanpied, T. A., Ehlers, M. D., and Weinberg, R. J. (2004). Lateral
10.025 organization of endocytic machinery in dendritic spines. Nat. Neurosci. 7,
Husi, H., Ward, M. A., Choudhary, J. S., Blackstock, W. P., and Grant, S. G. (2000). 917–918. doi: 10.1038/nn1303
Proteomic analysis of NMDA receptor-adhesion protein signaling complexes. Raghavachari, S., and Lisman, J. E. (2004). Properties of quantal transmission at
Nat. Neurosci. 3, 661–669. doi: 10.1038/76615 CA1 synapses. J. Neurophysiol. 92, 2456–2467. doi: 10.1152/jn.00258.2004
Inoue, A., and Okabe, S. (2003). The dynamic organization of postsynaptic Ryan, T. A., Myers, J., Holowka, D., Baird, B., and Webb, W. W. (1988). Molecular
proteins: translocating molecules regulate synaptic function. Curr. Opin. crowding on the cell surface. Science 239, 61–64. doi: 10.1126/science.
Neurobiol. 13, 332–340. doi: 10.1016/s0959-4388(03)00077-1 2962287
Kalashnikova, E., Lorca, R. A., Kaur, I., Barisone, G. A., Li, B., Ishimaru, T., Sainlos, M., Tigaret, C., Poujol, C., Olivier, N. B., Bard, L., Breillat, C.,
et al. (2010). SynDIG1: an activity-regulated, AMPA- receptor-interacting et al. (2011). Biomimetic divalent ligands for the acute disruption of
transmembrane protein that regulates excitatory synapse development. Neuron synaptic AMPAR stabilization. Nat. Chem. Biol. 7, 81–91. doi: 10.1038/nchem
65, 80–93. doi: 10.1016/j.neuron.2009.12.021 bio.498
Kerr, J. M., and Blanpied, T. A. (2012). Subsynaptic AMPA receptor distribution Santamaria, F., Gonzalez, J., Augustine, G. J., and Raghavachari, S. (2010).
is acutely regulated by actin-driven reorganization of the postsynaptic density. Quantifying the effects of elastic collisions and non-covalent binding on
J. Neurosci. 32, 658–673. doi: 10.1523/JNEUROSCI.2927-11.2012 glutamate receptor trafficking in the post-synaptic density. PLoS Comput. Biol.
Kusumi, A., Tsunoyama, T. A., Hirosawa, K. M., Kasai, R. S., and Fujiwara, T. K. 6:e1000780. doi: 10.1371/journal.pcbi.1000780
(2014). Tracking single molecules at work in living cells. Nat. Chem. Biol. 10, Santucci, D. M., and Raghavachari, S. (2008). The effects of NR2 subunit-
524–532. doi: 10.1038/nchembio.1558 dependent NMDA receptor kinetics on synaptic transmission and CaMKII
Legant, W. R., Shao, L., Grimm, J. B., Brown, T. A., Milkie, D. E., Avants, B. B., activation. PLoS Comput. Biol. 4:e1000208. doi: 10.1371/journal.pcbi.1000208
et al. (2016). High-density three-dimensional localization microscopy across Savin, T., and Doyle, P. S. (2005). Static and dynamic errors in particle
large volumes. Nat. Methods 13, 359–365. doi: 10.1038/nmeth.3797 tracking microrheology. Biophys. J. 88, 623–638. doi: 10.1529/biophysj.104.
Levet, F., Hosy, E., Kechkar, A., Butler, C., Beghin, A., Choquet, D., et al. (2015). 042457
SR-Tesseler: a method to segment and quantify localization-based super- Saxton, M. J. (1994). Anomalous diffusion due to obstacles: a monte carlo study.
resolution microscopy data. Nat. Methods 12, 1065–1071. doi: 10.1038/nmeth. Biophys. J. 66, 394–401. doi: 10.1016/s0006-3495(94)80789-1
3579 Schikorski, T., and Stevens, C. F. (1997). Quantitative ultrastructural analysis of
Li, T. P., Song, Y., MacGillavry, H. D., Blanpied, T. A., and Raghavachari, S. hippocampal excitatory synapses. J. Neurosci. 17, 5858–5867.
(2016). Protein crowding within the postsynaptic density can impede the escape Sheng, M., and Hoogenraad, C. C. (2007). The postsynaptic architecture of
of membrane proteins. J. Neurosci. 36, 4276–4295. doi: 10.1523/JNEUROSCI. excitatory synapses: a more quantitative view. Annu. Rev. Biochem. 76,
3154-15.2016 823–847. doi: 10.1146/annurev.biochem.76.060805.160029
Lu, H. E., MacGillavry, H. D., Frost, N. A., and Blanpied, T. A. (2014). Multiple Shinohara, Y., Hirase, H., Watanabe, M., Itakura, M., Takahashi, M., and
spatial and kinetic subpopulations of CaMKII in spines and dendrites as Shigemoto, R. (2008). Left-right asymmetry of the hippocampal synapses with
resolved by single-molecule tracking PALM. J. Neurosci. 34, 7600–7610. doi: 10. differential subunit allocation of glutamate receptors. Proc. Natl. Acad. Sci. U S
1523/JNEUROSCI.4364-13.2014 A 105, 19498–19503. doi: 10.1073/pnas.0807461105
MacGillavry, H. D., Kerr, J. M., and Blanpied, T. A. (2011). Lateral organization of Soto, D., Coombs, I. D., Renzi, M., Zonouzi, M., Farrant, M., and Cull-Candy, S. G.
the postsynaptic density. Mol. Cell. Neurosci. 48, 321–331. doi: 10.1016/j.mcn. (2009). Selective regulation of long-form calcium-permeable AMPA receptors
2011.09.001 by an atypical TARP, gamma-5. Nat. Neurosci. 12, 277–285. doi: 10.1038/nn.
MacGillavry, H. D., Kerr, J. M., Kassner, J., Frost, N. A., and Blanpied, T. A. (2016). 2266
Shank-cortactin interactions control actin dynamics to maintain flexibility of Spacek, J., and Harris, K. M. (1997). Three-dimensional organization of smooth
neuronal spines and synapses. Eur. J. Neurosci. 43, 179–193. doi: 10.1111/ejn. endoplasmic reticulum in hippocampal CA1 dendrites and dendritic spines of
13129 the immature and mature rat. J. Neurosci. 17, 190–203.
MacGillavry, H. D., Song, Y., Raghavachari, S., and Blanpied, T. A. (2013). Spacek, J., and Hartmann, M. (1983). Three-dimensional analysis of dendritic
Nanoscale scaffolding domains within the postsynaptic density concentrate spines. I. Quantitative observations related to dendritic spine and synaptic
synaptic AMPA receptors. Neuron 78, 615–622. doi: 10.1016/j.neuron.2013. morphology in cerebral and cerebellar cortices. Anat. Embryol. (Berl) 167,
03.009 289–310. doi: 10.1007/bf00298517
Manley, S., Gillette, J. M., Patterson, G. H., Shroff, H., Hess, H. F., Betzig, E., Steiner, P., Higley, M. J., Xu, W., Czervionke, B. L., Malenka, R. C., and
et al. (2008). High-density mapping of single-molecule trajectories with Sabatini, B. L. (2008). Destabilization of the postsynaptic density by PSD-95
photoactivated localization microscopy. Nat. Methods 5, 155–157. doi: 10. serine 73 phosphorylation inhibits spine growth and synaptic plasticity. Neuron
1038/nmeth.1176 60, 788–802. doi: 10.1016/j.neuron.2008.10.014
Nair, D., Hosy, E., Petersen, J. D., Constals, A., Giannone, G., Choquet, D., et al. Stewart, M. G., Medvedev, N. I., Popov, V. I., Schoepfer, R., Davies, H. A.,
(2013). Super-resolution imaging reveals that AMPA receptors inside synapses Murphy, K., et al. (2005). Chemically induced long-term potentiation increases
are dynamically organized in nanodomains regulated by PSD95. J. Neurosci. 33, the number of perforated and complex postsynaptic densities but does not
13204–13224. doi: 10.1523/JNEUROSCI.2381-12.2013 alter dendritic spine volume in CA1 of adult mouse hippocampal slices. Eur.
Okabe, S., Kim, H. D., Miwa, A., Kuriu, T., and Okado, H. (1999). Continual J. Neurosci. 21, 3368–3378. doi: 10.1111/j.1460-9568.2005.04174.x
remodeling of postsynaptic density and its regulation by synaptic activity. Nat. Tang, A.-H., Chen, H., Li, T. P., Metzbower, S. R., MacGillavry, H. D., and
Neurosci. 2, 804–811. doi: 10.1038/12175 Blanpied, T. A. (in press). A transsynaptic nanocolumn aligns neuotransmitter
Opazo, P., and Choquet, D. (2011). A three-step model for the synaptic release to receptors. Nature
recruitment of AMPA receptors. Mol. Cell. Neurosci. 46, 1–8. doi: 10.1016/j. Thompson, R. E., Larson, D. R., and Webb, W. W. (2002). Precise nanometer
mcn.2010.08.014 localization analysis for individual fluorescent probes. Biophys. J. 82,
Opazo, P., Sainlos, M., and Choquet, D. (2012). Regulation of AMPA receptor 2775–2783. doi: 10.1016/s0006-3495(02)75618-x
surface diffusion by PSD-95 slots. Curr. Opin. Neurobiol. 22, 453–460. doi: 10. Tomita, S., Chen, L., Kawasaki, Y., Petralia, R. S., Wenthold, R. J., Nicoll, R. A.,
1016/j.conb.2011.10.010 et al. (2003). Functional studies and distribution define a family of
Perez de Arce, K., Schrod, N., Metzbower, S. W., Allgeyer, E., Kong, G. K., transmembrane AMPA receptor regulatory proteins. J. Cell Biol. 161, 805–816.
Tang, A. H., et al. (2015). Topographic mapping of the synaptic cleft into doi: 10.1083/jcb.200212116
Uchida, N., Honjo, Y., Johnson, K. R., Wheelock, M. J., and Takeichi, M. postsynaptic receptor. Proc. Natl. Acad. Sci. U S A 94, 6983–6988. doi: 10.
(1996). The catenin/cadherin adhesion system is localized in synaptic junctions 1073/pnas.94.13.6983
bordering transmitter release zones. J. Cell Biol. 135, 767–779. doi: 10.1083/jcb. Zhang, J., and Diamond, J. S. (2009). Subunit- and pathway-specific localization
135.3.767 of NMDA receptors and scaffolding proteins at ganglion cell synapses
Valtschanoff, J. G., and Weinberg, R. J. (2001). Laminar organization of the in rat retina. J. Neurosci. 29, 4274–4286. doi: 10.1523/JNEUROSCI.5602-
NMDA receptor complex within the postsynaptic density. J. Neurosci. 21, 08.2009
1211–1217.
von Engelhardt, J., Mack, V., Sprengel, R., Kavenstock, N., Li, K. W., Stern- Conflict of Interest Statement: The authors declare that the research was
Bach, Y., et al. (2010). CKAMP44: a brain-specific protein attenuating short- conducted in the absence of any commercial or financial relationships that could
term synaptic plasticity in the dentate gyrus. Science 327, 1518–1522. doi: 10. be construed as a potential conflict of interest.
1126/science.1184178
Wang, Y., Cai, E., Rosenkranz, T., Ge, P., Teng, K. W., Lim, S. J., et al. Copyright © 2016 Li and Blanpied. This is an open-access article distributed
(2014). Small quantum dots conjugated to nanobodies as immunofluorescence under the terms of the Creative Commons Attribution License (CC BY). The use,
probes for nanometric microscopy. Bioconjug. Chem. 25, 2205–2211. doi: 10. distribution and reproduction in other forums is permitted, provided the original
1021/bc5004179 author(s) or licensor are credited and that the original publication in this journal
Xie, X., Liaw, J.-S., Baudry, M., and Berger, T. W. (1997). Novel expression is cited, in accordance with accepted academic practice. No use, distribution or
mechanism for synaptic potentiation: alignment of presynaptic release site and reproduction is permitted which does not comply with these terms.
The NMDA receptor (R) participates in many important physiological and pathological
processes. For example, its activation is required for both long-term potentiation (LTP)
and long-term depression (LTD) of synaptic transmission, cellular models of learning and
memory. Furthermore, it may play a role in the actions of amyloid-beta on synapses
as well as in the signaling leading to cell death following stroke. Until recently, these
processes were thought to be mediated by ion-flux through the receptor. Using a
combination of imaging and electrophysiological approaches, ion-flux independent
functions of the NMDAR were recently examined. In this review, we will discuss the
role of metabotropic NMDAR function in LTD and synaptic dysfunction.
Keywords: ion-flux independent, FRET-FLIM, long-term depression (LTD), NMDAR interactions with CaMKII,
amyloid-beta induced depression, PP1, excitotoxicity
INTRODUCTION
Transmembrane receptors have traditionally been divided into two classes: ionotropic and
Edited by: metabotropic. Ionotropic glutamate receptors (iGluRs) form channels that allow the passage of
George Augustine, ions into the cell to drive signaling, while metabotropic glutamate receptors (mGluRs) generate
Nanyang Technological University, downstream effects without ion-flux. The boundary between these two classes is not completely
Singapore distinct, as there has been evidence that several iGluRs are capable of producing effects in the
Reviewed by: absence of ion-flux. For example, the N-terminal domain of GluA2, a subunit of the AMPA
Kei Cho, receptor (AMPAR), is sufficient to promote spine formation in hippocampal neurons (Passafaro
University of Bristol, UK
et al., 2003). Another iGluR, the kainate receptor, can modulate GABA transmission without
William N. Green,
University of Chicago, USA ion-flux (Rodríguez-Moreno and Lerma, 1998).
The NMDA receptor (NMDAR), a member of the iGluR family, is ubiquitously expressed
*Correspondence:
Roberto Malinow
and plays numerous roles in the brain (Traynelis et al., 2010). Given its ability to conduct
[email protected] calcium ions (Ca2+ ) well, it has been assumed that downstream signaling triggered by NMDARs
was mediated by Ca2+ influx and increased cytoplasmic Ca2+ . However, to allow Ca2+ entry
Received: 25 April 2016 through the receptor, several conditions have to be fulfilled: (1) glutamate must bind to
Accepted: 06 July 2016 GluN2 subunits; (2) glycine, the co-agonist must bind to GluN1 subunits; and (3) neurons
Published: 28 July 2016 must be sufficiently depolarized to eliminate the voltage-dependent magnesium ion (Mg2+ )
Citation: block of the channel. During high-frequency stimulation (HFS) these three conditions are
Dore K, Aow J and Malinow R (2016) met resulting in long-term potentiation (LTP; Bliss and Collingridge, 1993). For long-term
The Emergence of NMDA Receptor
depression (LTD), however, the role of the NMDAR is not as clear. A long-standing model
Metabotropic Function: Insights
from Imaging.
has proposed that while LTP requires a large increase in cytoplasmic Ca2+ , a moderate rise
Front. Synaptic Neurosci. 8:20. in cytoplasmic Ca2+ would produce LTD (Lisman, 1989; Malenka, 1994). However, several
doi: 10.3389/fnsyn.2016.00020 recent studies indicate that NMDARs can induce LTD without ion-flow through the receptor
(Nabavi et al., 2013; Dore et al., 2015; Stein et al., 2015; Carter and during chemical LTD induction. This ligand-driven decrease
Jahr, 2016). Other publications have shown that excitotoxicity in FRET required NMDAR conformational movement but
as well as amyloid-beta-induced synaptic depression depend not PP1 activity (Aow et al., 2015). It is possible that
on NMDAR activity but are likewise independent of ion-flow the transient movement of PP1 relative to the NMDAR
(Kessels et al., 2013; Tamburri et al., 2013; Birnbaum et al., 2015; cytoplasmic domain exposes the catalytic active site of PP1
Weilinger et al., 2016). In this review, we will discuss how these to a target unavailable under basal conditions. One potential
studies probed NMDAR metabotropic activity with an emphasis target is calcium-calmodulin dependent protein kinase II
on the imaging techniques used. (CaMKII; Strack et al., 1997), which is recruited to the
NMDAR complex during LTP stimuli (Otmakhov et al.,
NMDAR-DEPENDENT LTD CAN BE 2004) and whose activity is required for both LTP (Malenka
INDUCED INDEPENDENTLY OF ION-FLUX et al., 1989; Malinow et al., 1989; Silva et al., 1992) and
LTD (Coultrap et al., 2014). By monitoring FRET between
Interestingly, evidence for ion-flux independent LTD can fluorescently-tagged GluN1 and CaMKII, a delayed decrease
be observed in data from older literature. Over 20 years in the NMDAR-CaMKII interaction was observed during ion-
ago, data were published indicating that MK-801, which flux independent LTD (Aow et al., 2015). This effect depended
blocks NMDAR channels, blocked LTP but failed to block on PP1 activity and was not seen with a CaMKII mutant
LTD (Mayford et al., 1995). A similar effect was obtained that cannot be dephosphorylated at Thr-286 (CaMKII-Thr-
by a different group (Scanziani et al., 1996). Surprisingly, 286-Asp), suggesting that dephosphorylation of Thr-286 is
these observations were not discussed in either study. The necessary to modify the NMDAR-CaMKII interaction (Aow
ion-flux dependence of LTD was recently examined more et al., 2015). Co-immunoprecipitation experiments additionally
closely (Nabavi et al., 2013). Low-frequency stimulation revealed that the amount of total CaMKII bound to the NMDAR
(LFS) produced LTD in the presence of either MK-801 or was unaffected by ion-flow independent LTD, whereas levels
7-chloro-kynurenate (7CK, a competitive GluN1 antagonist; of phosphorylated Thr-286 were reduced both during and
see Figure 1A) but not APV (a competitive GluN2 antagonist); after LTD induction (Aow et al., 2015). These results are
all three antagonists effectively blocked ion-flux through the consistent with a model for ion-flux independent LTD in which
NMDAR. Moreover, LTD was observed in experiments in which glutamate binding to the NMDAR induces a conformational
intracellular Ca2+ was clamped to basal levels, suggesting that change in the NMDAR intracellular domain that facilitates
a rise in intracellular Ca2+ is not required for LTD. It was PP1 access to and dephosphorylation of CaMKII at Thr-286,
thus proposed that glutamate binding to the NMDAR could thereby repositioning the CaMKII holoenzyme within the
induce a conformational change in the cytoplasmic domain NMDAR complex. The relocated CaMKII could in turn
of the NMDAR that triggers downstream signaling resulting potentially act on a novel site of the GluA1 subunit (Ser-567)
in LTD. that undergoes phosphorylation during LTD (Coultrap et al.,
To test if ligand binding could drive movement of the 2014). Consistent with this model, CaMKII phosphorylation of
NMDAR intracellular domain, FRET-FLIM [Forster resonance GluA1-Ser-567 does not require Ca2+ or calmodulin (Coultrap
energy transfer measured by fluorescence lifetime imaging of the et al., 2014). Ultimately, this process could increase AMPAR
FRET donor, see Toolbox and (Wallrabe and Periasamy, 2005; endocytosis (Lüscher et al., 1999; Lin et al., 2000; Kim
Yasuda, 2006)] was employed (Dore et al., 2015). Recombinant et al., 2001; Shi et al., 2001) and lead to depressed synaptic
GluN1 subunits of the NMDAR were tagged with GFP or transmission.
mCherry at their carboxyl(c)-terminus and co-expressed in Ion-flow independent NMDAR activation of downstream
neurons. As the magnitude of FRET is very sensitive to signaling pathways has also been linked to shrinkage of
the distance and orientation of the interacting fluorophores, dendritic spines (Stein et al., 2015). Stein et al. used 2-
nanometer-scale changes in distance can be reliably detected. photon laser scanning microscopy (TPLSM) to monitor
Bath application or uncaging of glutamate in the presence structural changes in the dendritic spines of GFP-
of MK-801 or 7CK, but not APV, produced a transient expressing hippocampal neurons. Low-frequency uncaging
change in FRET consistent with conformational movement of glutamate produced a ∼20% decrease in spine size
of the NMDAR cytoplasmic domain (Figure 1B). Infusing that was independent of ion-flow through the NMDAR
neurons with a GluN1 c-terminus antibody through a patch (Figure 1C). While high-frequency glutamate uncaging
pipette blocked the ligand-driven FRET change as well as LTD, produced an increase in spine volume, the same stimulus
suggesting that this conformational change is required for LTD in the presence of either 7CK or MK-801 led to spine
induction. shrinkage. This result is consistent with the finding that
Downstream signaling events were also examined using HFS (delivered electrically) in the presence of MK-801
FRET-FLIM (Aow et al., 2015). Protein-phosphatase 1 (PP1) produces LTD instead of LTP (Nabavi et al., 2013). Spine
is one of the first molecules whose activity was shown shrinkage was also abolished when p38 MAPK activity
to be required for LTD (Mulkey et al., 1993) and it was blocked (Stein et al., 2015), which is again consistent
co-immunoprecipitates with the NMDAR complex (Husi with the observation that levels of phosphorylated p38
et al., 2000). FRET between GluN1-GFP and PP1-mCherry, (which is the active form) were increased during ion-flow
observed in baseline conditions, was transiently reduced independent LTD (Nabavi et al., 2013). In the future, it will
be important to elucidate how the initial movement in the antagonists were typically acutely washed in and then out
NMDAR cytoplasmic domain subsequently affects signaling of the preparation during LTD induction (Volianskis et al.,
molecules, such as cofilin, calcineurin and p38, implicated 2015), instead of being present throughout the experimental
in LTD. duration. Furthermore, control pathways, which monitor
An ion-flux-independent mechanism for NMDAR- transmission onto the same neurons but do not receive the
dependent LTD has been challenged by some recent studies conditioning stimulus, were generally not included (Babiec
(Babiec et al., 2014; Volianskis et al., 2015; Sanderson et al., et al., 2014; Volianskis et al., 2015; Sanderson et al., 2016). These
2016) but confirmed by others (Kim et al., 2015; Stein differences in methodology are significant and can make an
et al., 2015; Carter and Jahr, 2016). It is notable that the impact in the outcome and interpretation of results (Nabavi
experimental conditions used in these recent studies that failed et al., 2014). Therefore, it will be important to compare carefully
to detect ion-flux-independent LTD were not identical to the experimental conditions employed by different studies.
those supporting this form of LTD. For instance, NMDAR Nevertheless, it remains possible that two different, independent
FRET can be measured by acquiring a series of images in different combinations of excitation and emission channels or by photobleaching of the FRET acceptor.
However, these methods are prone to errors and are generally not well suited for measurements in living cells expressing fluorescent proteins (Selvin, 2000;
Yasuda, 2006; Piston and Kremers, 2007). Fluorescence lifetime is defined as the average time a molecule stays in its excited state before emitting a photon (A).
Because lifetime is an intrinsic property of fluorophores, it is independent of experimental conditions such as concentration, excitation intensity and photobleaching
(Lakowicz, 2006; Yasuda, 2006). Importantly, by adding an additional route for the donor fluorophore to return to ground state, the degree of FRET makes the
fluorescence lifetime of the donor proportionally shorter (A,B). To measure fluorescence lifetimes, the most common approach is time correlated single photon
counting (TCSPC; Becker et al., 2004) which calculates the time delay between the detection of fluorescence photons and laser excitation pulses (B). When TCSPC is
combined with laser scanning microscopy, it is possible to obtain fluorescence lifetimes, and hence detect changes in donor-acceptor distances, at every pixel of an
image (B,C).
FRET-FLIM. (A) Jablonski diagram of FRET donor and acceptor energy levels. After excitation by a 1-photon (blue arrow) or 2-photon (brown arrows) laser,
the FRET donor can return to ground state by emitting a photon (green arrow) or by transferring its energy to a nearby acceptor (dashed green arrows). (B) The
fluorescence lifetime of the FRET donor, which becomes shorter with increased proximity of the FRET acceptor, is calculated with the fluorescence decay curve.
(C) Time-correlated single photon counting (TCSPC) records fluorescence decay curves for each pixel of an image. By fitting these curves, the fluorescence lifetime
of the FRET donor can be assessed. The FLIM image is then color coded according to the FRET donor lifetime at each pixel.
FIGURE 1 | Ion-flux independent long-term depression (LTD). (A) APV, but not 7-chloro-kynurenate (7CK), blocks LTD. LTD (1 Hz, black bar) was delivered to
the test pathway (black symbols), while the control pathway (red symbols) received no stimulus; the slice was incubated in APV (top) or 7CK (bottom) throughout the
experiment. Note that there was no change to the control or test pathway in APV, as expected, whereas the test pathway was significantly depressed after LTD
induction in 7CK; representative of N = 15, APV, p > 0.05 LTD; N = 15, 7CK, p < 0.01 LTD. Insets: superimposed traces obtained before and after LTD in the test
(black line) and control (red line) pathways. Scale bars 0.3 mV, 5 ms. Modified from Nabavi et al. (2013). (B) Chemical LTD induction (+NMDA) resulted in spines that
exhibit less Förster Resonance Energy Transfer (FRET; and higher GFP fluorescence lifetimes; ps = picoseconds) between the GFP- and mCherry-tagged GluN1
subunits (see Toolbox ), indicating movement of the NMDAR cytoplasmic domain. Modified from Dore et al. (2015). (C) A low frequency glutamate uncaging
protocol (LFU), similar to the 1 Hz electrical stimulation in (A), induced spine shrinkage both in vehicle or 7CK-incubated GFP-expressing neurons, indicating that
ion-flux through the NMDAR is not required for spine shrinkage. Modified from Stein et al. (2015).
forms of NMDAR-dependent LTD exist: one that requires of glycine was sufficient to prime NMDARs for subsequent
ion-flow through NMDARs and one that does not. Different use-dependent endocytosis, again leading to a decline in
experimental conditions could selectively recruit either of these NMDA currents (Nong et al., 2003). Moreover, synaptic
two forms. replacement of GluN2B- with GluN2A-containing NMDARs, an
important developmentally controlled process (Hestrin, 1992;
TRAFFICKING OF NMDAR IS REGULATED Monyer et al., 1994; Sheng et al., 1994; Stocca and Vicini,
BY SYNAPTIC ACTIVITY BUT NOT 1998; Tovar and Westbrook, 1999), required ligand binding
ION-FLUX without ion flux (Barria and Malinow, 2002). Interestingly,
the replacement of synaptic GluN2B-containing NMDARs by
In addition to its more recently described role in LTD, a newly synthetized GluN2B-containing NMDARs did not require
few older studies have indicated that ligand binding to the ligand binding. It is important to note that these effects
NMDAR, in the absence of ion-flux, could control NMDAR on NMDAR trafficking required both agonist and co-agonist
trafficking (Vissel et al., 2001; Barria and Malinow, 2002; binding to NMDARs as they were blocked by antagonists
Nong et al., 2003). The Westbrook lab showed that even to the glutamate binding site on GluN2 subunits or to the
with its pore blocked, ligand binding to the NMDAR drove glycine binding site on GluN1 subunits (Vissel et al., 2001;
tyrosine dephosphorylation of GluN2A subunits, resulting Barria and Malinow, 2002; Nong et al., 2003); in contrast, LTD
in NMDAR endocytosis and decreased NMDA currents. only requires ligand binding to GluN2 subunits (Nabavi et al.,
Another group separately observed that an initial application 2013).
FIGURE 2 | Metabotropic NMDAR activity can induce synaptic dysfunction. (A) APP-CT100 overexpression depresses synaptic transmission without ion-flux.
Model figure for a dual whole-cell recording of an infected and a neighboring uninfected CA1 neuron (top left) and an example trace of evoked AMPA receptor
(AMPAR) currents of such a recording (top right). Bottom, results from paired-recordings performed in ACSF containing no drug (N = 41), Ro 25-6981 (N = 37) or
MK-801 (N = 25); ∗∗ indicates p < 0.001. Modified from Kessels et al. (2013). (B) Spine density is reduced in a transgenic mouse model of Alzheimer’s disease (AD),
APV blocks the effect but not memantine. Scale bar, 5 µm. Modified from Birnbaum et al. (2015). (C) Sustained NMDA application causes excito-toxic insults to CA1
neurons in the form of blebbing of dendrites (left panels). This effect is blocked by APV and CGP-78608 (middle panels) but not MK-801 (right panels). Scale bar,
25 µm. Modified from Weilinger et al. (2016).
METABOTROPIC NMDAR ACTIVITY CAN shown using TPLSM or confocal microscopy that endogenously
INDUCE SYNAPTIC DYSFUNCTION expressed or exogenously applied amyloid-beta reduces spine
density in GFP-expressing neurons (Hsieh et al., 2006; Shrestha
Recent results have suggested a role for metabotropic NMDAR et al., 2006; Calabrese et al., 2007; Shankar et al., 2007; Wei
activity in amyloid-beta mediated synaptic dysfunction, which et al., 2010; Zempel et al., 2010), which may explain the
may contribute to hippocampal deficits in Alzheimer’s disease electrophysiologically observed decreased frequency of miniature
(AD) and precede neurological symptoms by a decade or excitatory postsynaptic currents (Kamenetz et al., 2003; Hsieh
more (Terry et al., 1991; Reiman et al., 1996). A number et al., 2006). Therefore, amyloid-beta induces synaptic insults
of studies using electrophysiology and imaging have reported that can be directly observed through imaging.
that amyloid-beta impairs LTP, depresses synaptic transmission A mechanism proposed to account for synaptic impairment
and induces synapse loss in various hippocampal preparations by amyloid-beta is enhanced ionotropic glutamate receptor
(Chapman et al., 1999; Larson et al., 1999; Walsh et al., 2002; endocytosis. Indeed, there is evidence that inhibiting endocytic
Wang et al., 2002; Kamenetz et al., 2003; Snyder et al., 2005; signaling pathways or overexpressing mutant endocytic-resistant
Hsieh et al., 2006; Shankar et al., 2007; Wei et al., 2010; receptors can ameliorate the reduction in AMPAR and/or
Birnbaum et al., 2015). The effect of intracellularly delivered NMDAR currents (Snyder et al., 2005; Hsieh et al., 2006; Knafo
amyloid-beta is not clear, as one report indicated synaptic et al., 2016). Notably, Kessels et al. (2013) reported that despite
depression (Ripoli et al., 2014) while another indicated synaptic block of ion flux, not all NMDAR antagonists prevented amyloid-
potentiation (Whitcomb et al., 2015). In many of these studies beta-induced depression of AMPAR-mediated transmission.
the electrophysiological results could be corroborated using The GluN2 antagonists (R)-CPP, Ro25-6981, and ifenprodil
imaging. For instance, Wei et al. (2010) used TPLSM to show afforded a complete block; whereas the GluN1 antagonist
that GFP-filled dendritic spines close to axons or dendrites 7CK and the NMDAR pore blocker MK-801 had no effect
overexpressing amyloid-beta displayed a smaller increase in (Figure 2A; Kessels et al., 2013). Thus, the block of depression
spine volume following a chemically-induced LTP protocol correlated with actions on different NMDAR subunits rather
as compared to more distant spines, suggesting that secreted than block of ion-flux. In another model, amyloid-beta
amyloid-beta impaired LTP. Hsieh et al. (2006) used TPLSM of oligomers exogenously applied to organotypic hippocampal
AMPARs tagged with the pH-sensitive GFP-variant SEP (Super- slices acutely depressed AMPAR-mediated transmission in
Ecliptic-Phluorin) to measure surface AMPARs and found that a manner that was dependent on synaptic stimulation and
amyloid-beta reduced surface GluA1 and GluA2. Likewise, NMDAR activation but not NMDAR ion-flux (Tamburri
immunostaining and imaging primary cultures treated with et al., 2013). Both studies therefore suggest that amyloid-
amyloid-beta revealed a reduction in surface NMDARs (Snyder beta activates a metabotropic NMDAR signaling pathway that
et al., 2005) and AMPARs (Almeida et al., 2005; Alfonso et al., depresses synaptic transmission. The evidence that this pathway
2014). The decrease in synaptic AMPAR and NMDAR currents could then be involved in eventual spine loss comes from
correlates, therefore, with a decrease in surface receptors as three other studies using imaging techniques. Two studies
determined with optical techniques. Finally, several groups have (Shankar et al., 2007, 2008) showed using TPLSM that (R)-CPP
FIGURE 3 | Model of NMDAR metabotropic functions. In baseline conditions, the NMDAR interacts with both protein-phosphatase 1 (PP1) and
calcium-calmodulin dependent protein kinase II (CaMKII) which is phosphorylated at Thr286. Upon ligand binding, movement in the NMDAR cytoplasmic domain
permits PP1 catalytic access to phospho-CaMKII-T286, this movement also inducing activation and phosphorylation of p38 MAPK. These signaling molecules (along
with others) will eventually lead to AMPAR removal. In the context of LTD, these events will produce spine shrinkage whereas in the case of sustained amyloid-beta
presence, spine elimination will occur.
prevented amyloid-beta-induced spine loss in GFP-expressing receptor is responsible for inducing cell death (Choi, 1995;
organotypic slice neurons. Birnbaum et al. (2015) subsequently Tu et al., 2010). Interestingly, recent findings suggest that a
demonstrated that the competitive GluN2 antagonist APV metabotropic NMDAR ‘‘signalsome’’—involving the NMDAR,
also blocked spine loss in transgenic AD mice (as well as the pannexin-1 channel (Panx1) and src kinase—is capable
in hippocampal slices incubated in amyloid-beta oligomers), of inducing cellular dysfunction in response to excessive
whilst MK-801, memantine, another NMDAR pore blocker, NMDAR stimulation (Weilinger et al., 2016). TPLSM was
and buffering postsynaptic calcium ions with BAPTA had used to image fluorescently labeled CA1 neurons in acute
no effect (Figure 2B). That study (Birnbaum et al., 2015) rat hippocampal slices treated with a high dose of NMDA.
also showed an amyloid-beta-induced reduction in PSD-95 This protocol induced blebbing in the dendrites of CA1
and synaptophysin levels that was blocked by APV but not neurons as well as mitochondrial dysfunction, an effect that
by MK-801 or memantine. Moreover, they demonstrated was blocked by co-application of the GluN1 antagonist CGP-
that p38 MAPK phosphorylation was increased by amyloid- 78608 and APV, but not by MK-801, indicating that ligand
beta in a NMDAR ion-flux independent manner and that binding to the NMDAR was capable of damaging neuronal
spine loss depended on p38 MAPK activity (Birnbaum et al., morphology independently of ion flux through the receptor
2015), supporting a link, previously examined (Hsieh et al., (Figure 2C). Interestingly, Panx1 channels appear to be
2006), between LTD and amyloid-beta-induced depression. located almost exclusively at the PSD (Zoidl et al., 2007),
Taken together, these imaging results are in agreement with which suggests that this metabotropic NMDAR ‘‘signalsome’’
electrophysiological experiments and support the hypothesis is synaptic. Additional experiments demonstrated that ligand
that amyloid-beta toxicity depresses synaptic transmission binding, but not NMDAR ion flux, was necessary for
via metabotropic NMDAR signaling that results in eventual downstream activation of Panx1 (Thompson et al., 2006, 2008;
spine loss. Weilinger et al., 2012). Metabotropic NMDAR activity did
The role of NMDARs in mediating excitotoxicity has been not change the degree of interaction between GluN1 and
extensively described (reviewed in Choi, 1992), and it has Panx1, but it did increase Src kinase binding to GluN1,
been widely suggested that excessive Ca2+ influx through the Src activation, as well as Src-dependent phosphorylation
of Panx1. Peptides that either disrupted the GluN1-Src mechanism is engaged during LTD. Ion-flux independent LTD
interaction (Src48 ) or interfered with Panx1 phosphorylation appears to be mediated by a movement in the NMDAR
(Tat-Panx308 ) were neuroprotective in vitro, and injection of cytoplasmic domain that affects its interactions with at least
Tat-Panx308 reduced brain lesion volume in an in vivo model two signaling proteins, PP1 and CaMKII. Subsequent signaling,
of stroke. Indeed, as ischemia in the brain is believed to drive with increased p38 MAPK phosphorylation likely, leads to
subsequent excitotoxicity, these results suggest that targeting AMPAR removal, shrinkage of dendritic spines and depressed
Src or Panx1 in a clinical setting could be therapeutically synaptic transmission (Figure 3). Interestingly, it seems that
effective. if the stimulus recruiting metabotropic NMDAR function is
sustained, as in the contexts of amyloid-beta overproduction
CONCLUDING REMARKS or excitotoxic conditions following ischemia, metabotropic
NMDAR activity can also lead to synaptic and neuronal
A number of studies have provided evidence that physiological dysfunction.
and pathological processes can be triggered by ligand binding
to the NMDAR, without requiring flow of ions through AUTHOR CONTRIBUTIONS
its pore. It will be important to determine conditions that
control whether an ionotropic or metabotropic NMDAR KD, JA and RM wrote the manuscript. KD designed figures.
REFERENCES Coultrap, S. J., Freund, R. K., O’Leary, H., Sanderson, J. L., Roche, K. W.,
Dell’Acqua, M. L., et al. (2014). Autonomous CaMKII mediates both LTP and
Alfonso, S., Kessels, H. W., Banos, C. C., Chan, T. R., Lin, E. T., Kumaravel, G., LTD using a mechanism for differential substrate site selection. Cell Rep. 6,
et al. (2014). Synapto-depressive effects of amyloid beta require PICK1. Eur. J. 431–437. doi: 10.1016/j.celrep.2014.01.005
Neurosci. 39, 1225–1233. doi: 10.1111/ejn.12499 Dore, K., Aow, J., and Malinow, R. (2015). Agonist binding to the NMDA receptor
Almeida, C. G., Tampellini, D., Takahashi, R. H., Greengard, P., Lin, M. T., drives movement of its cytoplasmic domain without ion flow. Proc. Natl. Acad.
Snyder, E. M., et al. (2005). Beta-amyloid accumulation in APP mutant neurons Sci. U S A 112, 14705–14710. doi: 10.1073/pnas.1520023112
reduces PSD-95 and GluR1 in synapses. Neurobiol. Dis. 20, 187–198. doi: 10. Gustiananda, M., Liggins, J. R., Cummins, P. L., and Gready, J. E. (2004).
1016/j.nbd.2005.02.008 Conformation of prion protein repeat peptides probed by FRET measurements
Aow, J., Dore, K., and Malinow, R. (2015). Conformational signaling required for and molecular dynamics simulations. Biophys. J. 86, 2467–2483. doi: 10.
synaptic plasticity by the NMDA receptor complex. Proc. Natl. Acad. Sci. U S A 1016/s0006-3495(04)74303-9
112, 14711–14716. doi: 10.1073/pnas.1520029112 Hestrin, S. (1992). Developmental regulation of NMDA receptor-mediated
Babiec, W. E., Guglietta, R., Jami, S. A., Morishita, W., Malenka, R. C., synaptic currents at a central synapse. Nature 357, 686–689. doi: 10.
and O’Dell, T. J. (2014). Ionotropic NMDA receptor signaling is required 1038/357686a0
for the induction of long-term depression in the mouse hippocampal Hsieh, H., Boehm, J., Sato, C., Iwatsubo, T., Tomita, T., Sisodia, S., et al.
CA1 region. J. Neurosci. 34, 5285–5290. doi: 10.1523/JNEUROSCI.5419- (2006). AMPAR removal underlies Aβ-induced synaptic depression and
13.2014 dendritic spine loss. Neuron 52, 831–843. doi: 10.1016/j.neuron.2006.
Barria, A., and Malinow, R. (2002). Subunit-specific NMDA receptor 10.035
trafficking to synapses. Neuron 35, 345–353. doi: 10.1016/s0896-6273(02) Husi, H., Ward, M. A., Choudhary, J. S., Blackstock, W. P., and Grant, S. G. (2000).
00776-6 Proteomic analysis of NMDA receptor-adhesion protein signaling complexes.
Becker, W., Bergmann, A., Hink, M. A., König, K., Benndorf, K., and Nat. Neurosci. 3, 661–669. doi: 10.1038/76615
Biskup, C. (2004). Fluorescence lifetime imaging by time-correlated Kamenetz, F., Tomita, T., Hsieh, H., Seabrook, G., Borchelt, D., Iwatsubo, T., et al.
single-photon counting. Microsc. Res. Tech. 63, 58–66. doi: 10.1002/jemt. (2003). APP processing and synaptic function. Neuron 37, 925–937. doi: 10.
10421 1016/s0896-6273(03)00124-7
Birnbaum, J. H., Bali, J., Rajendran, L., Nitsch, R. M., and Tackenberg, C. Kessels, H. W., Nabavi, S., and Malinow, R. (2013). Metabotropic NMDA receptor
(2015). Calcium flux-independent NMDA receptor activity is required for A function is required for β-amyloid-induced synaptic depression. Proc. Natl.
β oligomer-induced synaptic loss. Cell Death Dis. 6:e1791. doi: 10.1038/cddis. Acad. Sci. U S A 110, 4033–4038. doi: 10.1073/pnas.1219605110
2015.160 Kim, C. H., Chung, H. J., Lee, H. K., and Huganir, R. L. (2001). Interaction of the
Bliss, T. V., and Collingridge, G. L. (1993). A synaptic model of memory: long- AMPA receptor subunit GluR2/3 with PDZ domains regulates hippocampal
term potentiation in the hippocampus. Nature 361, 31–39. doi: 10.1038/36 long-term depression. Proc. Natl. Acad. Sci. U S A 98, 11725–11730. doi: 10.
1031a0 1073/pnas.211132798
Calabrese, B., Shaked, G. M., Tabarean, I. V., Braga, J., Koo, E. H., and Kim, Y., Hsu, C. L., Cembrowski, M. S., Mensh, B. D., and Spruston, N.
Halpain, S. (2007). Rapid, concurrent alterations in pre- and postsynaptic (2015). Dendritic sodium spikes are required for long-term potentiation at
structure induced by naturally-secreted amyloid-β protein. Mol. Cell. Neurosci. distal synapses on hippocampal pyramidal neurons. Elife 4:e06414. doi: 10.
35, 183–193. doi: 10.1016/j.mcn.2007.02.006 7554/eLife.06414
Carter, B. C., and Jahr, C. E. (2016). Postsynaptic, not presynaptic NMDA Knafo, S., Sánchez-Puelles, C., Palomer, E., Delgado, I., Draffin, J. E., Mingo, J.,
receptors are required for spike-timing-dependent LTD induction. Nature et al. (2016). PTEN recruitment controls synaptic and cognitive function in
Neuroscience. doi: 10.1038/nn.4343 [Epub ahead of print] Alzheimer’s models. Nat. Neurosci. 19, 443–453. doi: 10.1038/nn.4225
Chapman, P. F., White, G. L., Jones, M. W., Cooper-Blacketer, D., Marshall, V. J., Lakowicz, J. R. (2006). Principles of Fluorescence Spectroscopy. 3rd Edn. USA:
Irizarry, M., et al. (1999). Impaired synaptic plasticity and learning Spinger.
in aged amyloid precursor protein transgenic mice. Nat. Neurosci. 2, Larson, J., Lynch, G., Games, D., and Seubert, P. (1999). Alterations in synaptic
271–276. transmission and long-term potentiation in hippocampal slices from young and
Choi, D. W. (1992). Excitotoxic cell death. J. Neurobiol. 23, 1261–1276. doi: 10. aged PDAPP mice. Brain Res. 840, 23–35. doi: 10.1016/s0006-8993(99)01698-4
1002/neu.480230915 Lin, J. W., Ju, W., Foster, K., Lee, S. H., Ahmadian, G., Wyszynski, M., et al. (2000).
Choi, D. W. (1995). Calcium: still center-stage in hypoxic-ischemic neuronal Distinct molecular mechanisms and divergent endocytotic pathways of AMPA
death. Trends Neurosci. 18, 58–60. doi: 10.1016/0166-2236(95)80018-w receptor internalization. Nat. Neurosci. 3, 1282–1290. doi: 10.1038/81814
Lisman, J. (1989). A mechanism for the Hebb and the anti-Hebb processes Shankar, G. M., Bloodgood, B. L., Townsend, M., Walsh, D. M., Selkoe, D. J., and
underlying learning and memory. Proc. Natl. Acad. Sci. U S A 86, 9574–9578. Sabatini, B. L. (2007). Natural oligomers of the Alzheimer amyloid-β protein
doi: 10.1073/pnas.86.23.9574 induce reversible synapse loss by modulating an NMDA-type glutamate
Lüscher, C., Xia, H., Beattie, E. C., Carroll, R. C., von Zastrow, M., Malenka, R. C., receptor-dependent signaling pathway. J. Neurosci. 27, 2866–2875. doi: 10.
et al. (1999). Role of AMPA receptor cycling in synaptic transmission and 1523/JNEUROSCI.4970-06.2007
plasticity. Neuron 24, 649–658. doi: 10.1016/s0896-6273(00)81119-8 Shankar, G. M., Li, S., Mehta, T. H., Garcia-Munoz, A., Shepardson, N. E.,
Malenka, R. C. (1994). Synaptic plasticity in the hippocampus: LTP and LTD. Cell Smith, I., et al. (2008). Amyloid-β protein dimers isolated directly from
78, 535–538. doi: 10.1016/0092-8674(94)90517-7 Alzheimer’s brains impair synaptic plasticity and memory. Nat. Med. 14,
Malenka, R. C., Kauer, J. A., Perkel, D. J., Mauk, M. D., Kelly, P. T., Nicoll, R. A., 837–842. doi: 10.1038/nm1782
et al. (1989). An essential role for postsynaptic calmodulin and protein kinase Sheng, M., Cummings, J., Roldan, L. A., Jan, Y. N., and Jan, L. Y. (1994). Changing
activity in long-term potentiation. Nature 340, 554–557. doi: 10.1038/340554a0 subunit composition of heteromeric NMDA receptors during development of
Malinow, R., Schulman, H., and Tsien, R. W. (1989). Inhibition of postsynaptic rat cortex. Nature 368, 144–147. doi: 10.1038/368144a0
PKC or CaMKII blocks induction but not expression of LTP. Science 245, Shi, S., Hayashi, Y., Esteban, J. A., and Malinow, R. (2001). Subunit-specific
862–866. doi: 10.1126/science.2549638 rules governing AMPA receptor trafficking to synapses in hippocampal
Mayford, M., Wang, J., Kandel, E. R., and O’Dell, T. J. (1995). CaMKII pyramidal neurons. Cell 105, 331–343. doi: 10.1016/s0092-8674(01)
regulates the frequency-response function of hippocampal synapses for the 00321-x
production of both LTD and LTP. Cell 81, 891–904. doi: 10.1016/0092-8674(95) Shrestha, B. R., Vitolo, O. V., Joshi, P., Lordkipanidze, T., Shelanski, M., and
90009-8 Dunaevsky, A. (2006). Amyloid β peptide adversely affects spine number and
Monyer, H., Burnashev, N., Laurie, D. J., Sakmann, B., and Seeburg, P. H. motility in hippocampal neurons. Mol. Cell. Neurosci. 33, 274–282. doi: 10.
(1994). Developmental and regional expression in the rat brain and functional 1016/j.mcn.2006.07.011
properties of four NMDA receptors. Neuron 12, 529–540. doi: 10.1016/0896- Silva, A. J., Stevens, C. F., Tonegawa, S., and Wang, Y. (1992). Deficient
6273(94)90210-0 hippocampal long-term potentiation in alpha-calcium-calmodulin kinase II
Mulkey, R. M., Herron, C. E., and Malenka, R. C. (1993). An essential role mutant mice. Science 257, 201–206. doi: 10.1126/science.1378648
for protein phosphatases in hippocampal long-term depression. Science 261, Snyder, E. M., Nong, Y., Almeida, C. G., Paul, S., Moran, T., Choi, E. Y., et al.
1051–1055. doi: 10.1126/science.8394601 (2005). Regulation of NMDA receptor trafficking by amyloid-β. Nat. Neurosci.
Nabavi, S., Fox, R., Alfonso, S., Aow, J., and Malinow, R. (2014). GluA1 trafficking 8, 1051–1058. doi: 10.1038/nn1503
and metabotropic NMDA: addressing results from other laboratories Stein, I. S., Gray, J. A., and Zito, K. (2015). Non-ionotropic NMDA receptor
inconsistent with ours. Philos. Trans. R. Soc. Lond. B Biol. Sci. 369:20130145. signaling drives activity-induced dendritic spine shrinkage. J. Neurosci. 35,
doi: 10.1098/rstb.2013.0145 12303–12308. doi: 10.1523/JNEUROSCI.4289-14.2015
Nabavi, S., Kessels, H. W., Alfonso, S., Aow, J., Fox, R., and Malinow, R. (2013). Stocca, G., and Vicini, S. (1998). Increased contribution of NR2A subunit to
Metabotropic NMDA receptor function is required for NMDA receptor- synaptic NMDA receptors in developing rat cortical neurons. J. Physiol. 507,
dependent long-term depression. Proc. Natl. Acad. Sci. U S A 110, 4027–4032. 13–24. doi: 10.1111/j.1469-7793.1998.013bu.x
doi: 10.1073/pnas.1219454110 Strack, S., Barban, M. A., Wadzinski, B. E., and Colbran, R. J. (1997). Differential
Nong, Y., Huang, Y. Q., Ju, W., Kalia, L. V., Ahmadian, G., Wang, Y. T., et al. inactivation of postsynaptic density-associated and soluble Ca2+ /calmodulin-
(2003). Glycine binding primes NMDA receptor internalization. Nature 422, dependent protein kinase II by protein phosphatases 1 and 2A. J. Neurochem.
302–307. doi: 10.1038/nature01497 68, 2119–2128. doi: 10.1046/j.1471-4159.1997.68052119.x
Otmakhov, N., Tao-Cheng, J. H., Carpenter, S., Asrican, B., Dosemeci, A., Tamburri, A., Dudilot, A., Licea, S., Bourgeois, C., and Boehm, J. (2013).
Reese, T. S., et al. (2004). Persistent accumulation of calcium/calmodulin- NMDA-receptor activation but not ion flux is required for amyloid-β
dependent protein kinase II in dendritic spines after induction of NMDA induced synaptic depression. PloS One 8:e65350. doi: 10.1371/journal.pone.
receptor-dependent chemical long-term potentiation. J. Neurosci. 24, 0065350
9324–9331. doi: 10.1523/JNEUROSCI.2350-04.2004 Terry, R. D., Masliah, E., Salmon, D. P., Butters, N., DeTeresa, R., Hill, R., et al.
Passafaro, M., Nakagawa, T., Sala, C., and Sheng, M. (2003). Induction of dendritic (1991). Physical basis of cognitive alterations in Alzheimer’s disease: synapse
spines by an extracellular domain of AMPA receptor subunit GluR2. Nature loss is the major correlate of cognitive impairment. Ann. Neurol. 30, 572–580.
424, 677–681. doi: 10.1038/nature01781 doi: 10.1002/ana.410300410
Piston, D. W., and Kremers, G. J. (2007). Fluorescent protein FRET: the good, the Thompson, R. J., Jackson, M. F., Olah, M. E., Rungta, R. L., Hines, D. J.,
bad and the ugly. Trends Biochem. Sci. 32, 407–414. doi: 10.1016/j.tibs.2007. Beazely, M. A., et al. (2008). Activation of pannexin-1 hemichannels augments
08.003 aberrant bursting in the hippocampus. Science 322, 1555–1559. doi: 10.
Reiman, E. M., Caselli, R. J., Yun, L. S., Chen, K., Bandy, D., Minoshima, S., et al. 1126/science.1165209
(1996). Preclinical evidence of Alzheimer’s disease in persons homozygous for Thompson, R. J., Zhou, N., and MacVicar, B. A. (2006). Ischemia opens neuronal
the epsilon 4 allele for apolipoprotein E. N. Engl. J. Med. 334, 752–758. doi: 10. gap junction hemichannels. Science 312, 924–927. doi: 10.1126/science.
1056/nejm199603213341202 1126241
Ripoli, C., Cocco, S., Li Puma, D. D., Piacentini, R., Mastrodonato, A., Scala, F., Tovar, K. R., and Westbrook, G. L. (1999). The incorporation of NMDA receptors
et al. (2014). Intracellular accumulation of amyloid-β (Aβ) protein plays a with a distinct subunit composition at nascent hippocampal synapses in vitro.
major role in Aβ-induced alterations of glutamatergic synaptic transmission J. Neurosci. 19, 4180–4188.
and plasticity. J. Neurosci. 34, 12893–12903. doi: 10.1523/JNEUROSCI.1201- Traynelis, S. F., Wollmuth, L. P., McBain, C. J., Menniti, F. S., Vance, K. M.,
14.2014 Ogden, K. K., et al. (2010). Glutamate receptor ion channels: structure,
Rodríguez-Moreno, A., and Lerma, J. (1998). Kainate receptor modulation of regulation and function. Pharmacol. Rev. 62, 405–496. doi: 10.1124/pr.109.
GABA release involves a metabotropic function. Neuron 20, 1211–1218. doi: 10. 002451
1016/s0896-6273(00)80501-2 Tu, W., Xu, X., Peng, L., Zhong, X., Zhang, W., Soundarapandian, M. M., et al.
Sanderson, J. L., Gorski, J. A., and Dell’Acqua, M. L. (2016). NMDA receptor- (2010). DAPK1 interaction with NMDA receptor NR2B subunits mediates
dependent LTD requires transient synaptic incorporation of Ca2+ -permeable brain damage in stroke. Cell 140, 222–234. doi: 10.1016/j.cell.2009.12.055
AMPARs mediated by AKAP150-anchored PKA and calcineurin. Neuron 89, Vissel, B., Krupp, J. J., Heinemann, S. F., and Westbrook, G. L. (2001). A use-
1000–1015. doi: 10.1016/j.neuron.2016.01.043 dependent tyrosine dephosphorylation of NMDA receptors is independent of
Scanziani, M., Malenka, R. C., and Nicoll, R. A. (1996). Role of intercellular ion flux. Nat. Neurosci. 4, 587–596. doi: 10.1038/88404
interactions in heterosynaptic long-term depression. Nature 380, 446–450. Volianskis, A., France, G., Jensen, M. S., Bortolotto, Z. A., Jane, D. E., and
doi: 10.1038/380446a0 Collingridge, G. L. (2015). Long-term potentiation and the role of N-methyl-
Selvin, P. R. (2000). The renaissance of fluorescence resonance energy transfer. D-aspartate receptors. Brain Res. 1621, 5–16. doi: 10.1016/j.brainres.2015.
Nat. Struct. Biol. 7, 730–734. doi: 10.1038/78948 01.016
Wallrabe, H., and Periasamy, A. (2005). Imaging protein molecules using FRET Yasuda, R. (2006). Imaging spatiotemporal dynamics of neuronal signaling using
and FLIM microscopy. Curr. Opin. Biotechnol. 16, 19–27. doi: 10.1016/j.copbio. fluorescence resonance energy transfer and fluorescence lifetime imaging
2004.12.002 microscopy. Curr. Opin. Neurobiol. 16, 551–561. doi: 10.1016/j.conb.2006.
Walsh, D. M., Klyubin, I., Fadeeva, J. V., Cullen, W. K., Anwyl, R., Wolfe, M. S., 08.012
et al. (2002). Naturally secreted oligomers of amyloid β protein potently inhibit Zempel, H., Thies, E., Mandelkow, E., and Mandelkow, E. M. (2010).
hippocampal long-term potentiation in vivo. Nature 416, 535–539. doi: 10. Aβ oligomers cause localized Ca2+ elevation, missorting of endogenous
1038/416535a Tau into dendrites, Tau phosphorylation and destruction of microtubules
Wang, H. W., Pasternak, J. F., Kuo, H., Ristic, H., Lambert, M. P., Chromy, B., et al. and spines. J. Neurosci. 30, 11938–11950. doi: 10.1523/JNEUROSCI.2357-
(2002). Soluble oligomers of β amyloid (1–42) inhibit long-term potentiation 10.2010
but not long-term depression in rat dentate gyrus. Brain Res. 924, 133–140. Zoidl, G., Petrasch-Parwez, E., Ray, A., Meier, C., Bunse, S., Habbes, H. W.,
doi: 10.1016/s0006-8993(01)03058-x et al. (2007). Localization of the pannexin1 protein at postsynaptic sites in
Wei, W., Nguyen, L. N., Kessels, H. W., Hagiwara, H., Sisodia, S., and Malinow, R. the cerebral cortex and hippocampus. Neuroscience 146, 9–16. doi: 10.1016/j.
(2010). Amyloid β from axons and dendrites reduces local spine number and neuroscience.2007.01.061
plasticity. Nat. Neurosci. 13, 190–196. doi: 10.1038/nn.2476
Weilinger, N. L., Lohman, A. W., Rakai, B. D., Ma, E. M., Bialecki, J., Conflict of Interest Statement: The authors declare that the research was
Maslieieva, V., et al. (2016). Metabotropic NMDA receptor signaling couples conducted in the absence of any commercial or financial relationships that could
Src family kinases to pannexin-1 during excitotoxicity. Nat. Neurosci. 19, be construed as a potential conflict of interest.
432–442. doi: 10.1038/nn.4236
Weilinger, N. L., Tang, P. L., and Thompson, R. J. (2012). Anoxia-induced NMDA Copyright © 2016 Dore, Aow and Malinow. This is an open-access article distributed
receptor activation opens pannexin channels via Src family kinases. J. Neurosci. under the terms of the Creative Commons Attribution License (CC BY). The use,
32, 12579–12588. doi: 10.1523/jneurosci.1267-12.2012 distribution and reproduction in other forums is permitted, provided the original
Whitcomb, D. J., Hogg, E. L., Regan, P., Piers, T., Narayan, P., Whitehead, G., et al. author(s) or licensor are credited and that the original publication in this journal
(2015). Intracellular oligomeric amyloid-β rapidly regulates GluA1 subunit of is cited, in accordance with accepted academic practice. No use, distribution or
AMPA receptor in the hippocampus. Sci. Rep. 5:10934. doi: 10.1038/srep10934 reproduction is permitted which does not comply with these terms.
The emergence of such dual approaches, termed correlative heavy elements (i.e., high atomic number) in the uncoated sample
light electron microscopy (CLEM), from a sparsely known can also be used to render images comparable to transmission
branch of imaging to center stage can be linked to a series electron micrographs of ultrathin sections, an approach that is
of landmark papers from the nineties (Deerinck et al., 1994; of particular relevance for embedded sections on an electron-
Svitkina et al., 1995). Since then, fueled by advances in imaging opaque surface or samples that are investigated ‘en block’
and labeling techniques, publications using CLEM approaches (Briggman et al., 2011).
have steadily been rising. In this review, we will revisit what A single cross-section, or the surface of the sample, is lacking
developments have incited the advent of correlative imaging information on the three-dimensional organization within the
approaches, and highlight the most relevant dual microscopy biological specimen. Here, cryo-fracture, where frozen samples
techniques that are currently available. We begin by giving an are broken to expose cell structures along the fracture line
overview of electron and light-based imaging methods that can be (Castejon and Caraballo, 1980; Castejon, 1996), or unroofing,
used for CLEM, followed by a section describing how individual where intracellular structures are uncovered by brief bursts of
approaches can be successfully combined to create correlative ultrasound (Lang, 2003), can be used together with deep-etching
approaches. Next, we will reflect on technical concerns that need (Heuser and Salpeter, 1979) to gain insights into subcellular
to be taken care of when designing dual imaging experiments, organization. While useful, these strategies lack the capability for
before closing with a discussion on current limitations and future systematic three-dimensional sample reconstruction that can be
challenges associated with CLEM. achieved with serial sectioning, where successive ultrathin slices
are imaged. Early attempts of serial sectioning reach back to
the late 60s (De Rosier and Klug, 1968), and have since then
ELECTRON AND LIGHT-BASED continuously been used to investigate neuronal networks (White
TECHNIQUES USED FOR CLEM et al., 1986) and synaptic connection (Mishchenko et al., 2010).
In modern serial section TEM (ssTEM), sections of 60 ± 20 nm
CLEM describes a continuously growing number of procedures are cut with an ultra-microtome and placed manually on a metal
that allow merging electron and light-based images from the support grid (Harris et al., 2006). More recently, automated tape-
same object. Thus, at least in theory, each existing electron and collecting ultra-microtome SEM (ATUM-SEM) has emerged as a
light microscopy technique where ultrastructure remains intact powerful alternative (Hayworth et al., 2006; Kasthuri et al., 2007,
could be paired to generate a CLEM image. To illustrate this 2015; Terasaki et al., 2013; Morgan et al., 2016). Here, sections
combinatorial potential, and to expound existing limitations, we of 30 nm are cut by an ultra-microtome and collected from the
begin this section by discussing electron and light microscopy water bath using a conveyor-belt like support tape. As the support
techniques suitable for correlative approaches, before proceeding tape is electron-opaque, a finely focused SEM beam is applied
to examples where CLEM was successful applied to study to raster the surface of the sample and backscattered electrons
neuronal and synaptic function. are used to reconstruct the image. Thus, compared to ssTEM,
ATUM-based strategies not only allow thinner sectioning and
Electron Microscopy–Visualizing the rapid imaging of larger areas but also substantially reduce errors
Ultrastructure associated with manual sample handling (Hayworth et al., 2014).
Electron micrographs in CLEM rely to a large extent on Alternatively, an embedded tissue block can also be sectioned
transmission electron microscopy (TEM) and scanning electron directly within the SEM vacuum chamber, either using a diamond
microscopy (SEM). TEM, which enables visualization of 50– knife (serial block-face SEM, SBEM; Leighton, 1981; Denk and
100 nm thick cross-sections of samples with a resolution of Horstmann, 2004) or by milling with a focused ion-beam (FIB-
down to a few Angstrom (Pierce and Buseck, 1974; Ruska, SEM; Watkins et al., 1986; Knott et al., 2008). In the latter, a
1987; Erni et al., 2009), was critically involved in the detailed scanning electron microscope is used to image the surface of
characterization of synaptic structures (Gray, 1959; Pappas and the embedded sample, while a high current focused ion beam
Bennett, 1966; Fifkova and Delay, 1982; Landis and Reese, 1983; continuously slices off sections perpendicular to the SEM axis,
Watanabe et al., 2013, 2014), and has helped to advance our thus allowing 3D reconstruction of the sample. Compared to
understanding on age and disease-dependent changes in synaptic diamond-based sectioning, FIB-SEM is thus not only faster but
properties (DeKosky and Scheff, 1990; Navlakha et al., 2013). also permits sectioning samples with a step size in the single nm
Contrary to TEM, where electron shadows are used to range (Knott et al., 2008; Merchan-Perez et al., 2009; Lehmann
create the image, SEM-based strategies utilizes the interaction et al., 2014).
of electrons with molecules in the sample to recreate the Another strategy for detailed three-dimensional
image. One commonly applied SEM strategy, where secondary reconstructions relies on electron tomography (ET). In this
electrons are used to determine the surface topography of objects approach, a tilt series (usually ranging from −60◦ to +60◦ )
with a precision of ∼1 nm (Vernon-Parry, 2000), was central of two-dimensional images is generated to reconstruct the
for investigating surface features such as the morphology of three-dimensional shape of an object within the single slice
neuromuscular junctions (Desaki and Uehara, 1981) as well as (Crowther et al., 1970; Hoppe et al., 1974; Penczek, 2010; Ercius
structural reorganization during exploratory dendritic filopodia et al., 2015). ET was successfully used to resolve ultrastructural
formation in hippocampal neurons (Galic et al., 2014). In features of the presynapse (Perkins et al., 2015) and synaptic
addition, back-scattered electrons created by interaction with vesicle populations in saccular hair cells (Lenzi et al., 1999). Yet,
while ET is suitable for atomic-resolution, radiation damage, among others subcellular protein localization (Chalfie et al., 1994;
and limitations in interpretability for thicker sections need to be Heim et al., 1994; Kilgore et al., 2013), protein activity (Adams
considered when preparing the sample (Tocheva et al., 2010). et al., 1991), ion dynamics (Tsien, 1980, 1981; Tsien et al., 1982;
Finally, we would like to note that further electron-based Minta and Tsien, 1989; Nakai et al., 2001), pH (Tanasugarn
imaging strategies exist (e.g., STEM (Crewe et al., 1970; Engel, et al., 1984; Miesenbock et al., 1998), membrane potential or
1978), tSEM (Kuwajima et al., 2013)). While not commonly used voltage (Davila et al., 1973; Siegel and Isacoff, 1997; Zochowski
in CLEM studies, it is plausible to envision that some of these et al., 2000), or lipid species (Stauffer et al., 1998). Intriguingly,
additional types of electron microscopy may be beneficial for light is also suitable to actively manipulate the biological sample
a particular question. For readers interested in learning more in a precise spatio-temporal manner. One commonly used
on electron microscopy techniques, we recommend reading one approach relies on light-activation of caged substrates. Relevant
of the many excellent reviews available on this exciting topic for neurobiology, strategies for uncaging calcium (Ellisdavies and
(Briggman and Bock, 2012; Knott and Genoud, 2013; Miranda Kaplan, 1994), IP3 (Wang and Augustine, 1995), and various
et al., 2015; Perkins et al., 2015). neurotransmitters (Ellis-Davies, 2011) have been realized. More
recently, light has also been used to regulate protein function
(Cao et al., 2008; Tischer and Weiner, 2014). This approach,
Light Microscopy–Monitoring and coined optogenetics, has provided among others tools for
Manipulating Cell Function controlling neuronal and synaptic activity with unprecedented
While the electron microscopy techniques described above are spatio-temporal precision (Boyden et al., 2005; Fenno et al.,
well suited to investigate ultrastructural features with high 2011; Rost et al., 2015). Finally, it is worth mentioning that even
axial resolution, they lack spatio-temporal information available physical cell parameters, such as shape or membrane tension, can
with light microscopy. To increase penetration depth and be altered by light in living samples, using for instance optical
reduce background compared to whole field illumination (Epi), tweezers (Ashkin, 1970; Zhang and Liu, 2008). In summary,
CLEM studies frequently rely on confocal microscopy, where when combined with electron microscopy, these light-based
a pinhole is used to reduce out-of-focus light (Minsky, 1988), approaches allow precisely altering a specific parameter while
light sheet microscopy, where a light beam perpendicular to monitoring the cellular responses followed by analysis of the
the objective illuminates only the focal plane (Engelbrecht and corresponding ultrastructural features.
Stelzer, 2006; Keller et al., 2008), or two-photon microscopy,
where simultaneous absorption of two photons is used for
spatially controlled illumination (Denk et al., 1990). Resolution Correlative Light and Electron
in light and electron microscopy is based on numerical aperture Microscopy in Neuroscience
and wavelength. However, unlike in electron microscopy, where Although the potential of combining ultrastructural information
resolution is limited by the spot size (SEM) or the grain of the with functional studies was early noticed, first attempts to
detector (TEM), the limiting factors in light microscopy is the combine these microscopy techniques were limited to depicting
wavelength (Rayleigh, 1879; Abbe, 1883). To reach beyond the separately prepared and imaged biological samples (Porter et al.,
diffraction limit, several approaches have been introduced over 1945). Experiments where fluorescence and EM images of
the last decade. Among the most prominent super-resolution subcellular structures from the same cell were aligned started
microscopy techniques used in CLEM studies are stimulated appearing in the 1970s (Hollander, 1970; Nakai and Iwashita,
emission depletion microscopy (STED), where a depletion laser 1976), and have since then been applied to investigate a variety
limits the width of the emitting light source (Hell and Wichmann, of neurobiological questions.
1994; Klar et al., 2001), and stochastic techniques, such as To fully understand neuronal or synaptic function, the cellular
fluorescent photo-activation localization microscopy (PALM; context needs to be considered. It is thus not surprising, that large
Betzig and Chichester, 1993; Betzig et al., 2006) or stochastic efforts have been put into elucidating ultrastructural information
optical reconstruction microscopy (STORM; Rust et al., 2006), within the intact tissue. In situ CLEM, where data from life-
where light emitted from sequentially activated fluorophores is cell fluorescence imaging is combined with EM micrographs,
fitted to determine its precise localization. As before, we would has advanced our understanding of neuronal circuits (Bock
like to note that additional light-based approaches [e.g., total et al., 2011; Briggman et al., 2011; Lee et al., 2016), and
internal reflection fluorescence microscopy (Axelrod, 1981) or has also shed light on subcellular behaviors such as axosome
structured illumination microscopy (Neil et al., 1997)] can be shedding (Figure 1A) (Bishop et al., 2004). An alternative
used in CLEM approaches. Readers interested to learn more strategy, frequently applied to gain information of individual
about fluorescence microscopy techniques, we refer to one of the neurons in situ, relies on array tomography (AT), where plastic-
many excellent reviews focused on these topics (Lichtman and embedded tissue samples are sliced with an ultra-microtome,
Conchello, 2005; Combs, 2010; Huang et al., 2010). bonded array-wise onto a glass coverslip, stained, and finally
Fluorescence light microscopy not only allows to visualize cells imaged by fluorescence and electron microscopy (Figure 1B)
within neuronal circuits (Livet et al., 2007), but a continuously (Micheva and Smith, 2007; Rah et al., 2013; Collman et al.,
growing number of molecular probes and genetically encoded 2015). By combining super-resolution techniques and serial
markers also provide tools to study a variety of neuronal and sectioning, it was even possible to detect presynaptic dense
synaptic properties. Parameters that can be analyzed include projection proteins at a lateral resolution of 35–65 nm (Watanabe
et al., 2011). While serial sectioning-based CLEM approaches penetration of the antibody. In the post-embedding method,
allow studying biological samples with high axial and lateral the antigen–antibody reaction is performed after (i.e., post)
resolution, sample preparation, and 3D image alignment can be plastic embedding. As labeling takes place on thin tissue slices,
challenging (Micheva and Smith, 2007; Collman et al., 2015). antigens are more easily accessible. However, OsO4 , which is
Here, SBEM and FIB-SEM based approaches, where en-block often used as fixative and stain for membrane structures, can
EM imaging and fluorescence approaches are merged to obtain quench the fluorescence signal, and epoxy-based resins used
CLEM images, have helped to streamline in situ studies focused due to its good preservation and ultra-sectioning properties
on the complexity of the nervous system (Briggman et al., 2011; partially inhibit photo-switching of fluorophores (Kim et al.,
Blazquez-Llorca et al., 2015), or to reconstruct whole cortical 2015). Here, a number of modifications, such as using acrylic
neurons (Maco et al., 2013) and single synapses (Bosch et al., resins as embedding material or replacing OsO4 with uranyl
2015). acetate, tannic acid or p-phenylenediamine, have helped to
Orthogonal to in situ approaches described above, CLEM enhance compatibility and immunoreactivity (Phend et al.,
can also be applied to neurons isolated from brain tissues and 1995; Osamura et al., 2000; Akagi et al., 2006; Kim et al.,
cultured in vitro (Banker and Cowan, 1977). As such, cultured 2015). Complementing these slice-based assays, cells can also be
neurons are compatible with each CLEM technique described prepared for analysis of surface features. As before, samples are
above (Figure 1C) (Al Jord et al., 2014; Paez-Segala et al., 2015; fixed. However, rather than embedding the sample, a platinum
Fares et al., 2016). Although these cells lack the physiological replica of the surface is generated (Heuser et al., 1976), or the
context usually encountered by neurons within a tissue, cultured sample is subjected to critical point drying and surface staining,
neurons may still yield advantages compared to in situ samples prior to image acquisition via SEM. Finally, one last concern
depending on the biological question. One such example, and is that chemical fixation is not instantaneous (Szczesny et al.,
complementing the section-based CLEM assays described so 1996). To circumvent this issue, quick-freezing, where samples
far, is the analysis of processes at the cellular surface, such as are ‘slammed’ on a super-cold block of metal sprayed with
curvature-dependent protein recruitment to deforming plasma liquid helium (Heuser et al., 1976) or high-pressure freezing,
membranes (Figure 1D) (Galic et al., 2014). where samples are frozen in milli seconds while pressure
is increased to avoid water crystallization (Hunziker et al.,
1984; Dahl and Staehelin, 1989), have improved preservation
TECHNICAL CONSIDERATIONS quality.
In this section, we will focus on technical possibilities and Picking the Right Markers for CLEM
limitations that need to be taken into consideration when Unfortunately, not all markers are equally well suited for
designing a CLEM experiment. We start by discussing challenges individual CLEM approaches. To account for these limitations,
arising from sample preparation. Next, we will survey what a series of labeling techniques has emerged. Considering its
markers are suitable for which technique, and discuss strategies widespread availability, the most commonly used markers for
for alignment of corresponding fluorescence images and electron CLEM are genetically encoded fluorescence tags, such as GFP,
micrographs with respect to their potential and drawbacks. that can be detected in electron micrographs via immuno-
gold labeling (Figure 2A) (Kukulski et al., 2011; Galic et al.,
Sample Preparation for CLEM Images 2012). In order to distinguish several proteins in the same
The quality of correlative image alignment critically relies on sample, immuno-gold particles of variable sizes directed against
the ability to maintain the native organization of the cell different fluorescence tags can be used. However, since sample
during fixation and subsequent sample preparation. Thus, CLEM fixation may interfere with genetically encoded fluorescence
techniques not only need to be optimized for signal strength, but markers (Muller-Reichert and Verkade, 2014) and gold particles
also for shape preservation. Formaldehyde, glutaraldehyde, and may quench the fluorophore in parallel fluorescence/electron
osmium tetroxide (OsO4 ), have been the standard fixatives for imaging (Kandela and Albrecht, 2007), additional markers are
decades, including for brain tissues that are due to their softness desirable. One such alternative relies on quantum dots (QDs;
easily damaged during fixation and subsequent processing. Figure 2B), which were shown to have a 10 times higher labeling
Classical fixation protocols include whole animal perfusion via efficiency than immuno-gold (Giepmans et al., 2005; Kuipers
the vascular system for larger specimens (e.g., whole brain) et al., 2015). These semiconductor nanocrystals consist of a
as well as immersion of tissue slice and cultured cells. Once cadmium selenide core surrounded by a zinc sulfide shell coated
the biological sample is fixed, the actual preparation for the with affinity ligands (e.g., antibodies) for targeting the desired
respective EM technique is rendered. For slice-based assays, biomolecules (Alivisatos, 1996). Intriguingly, as the fluorescence
two different approaches exist: the pre-embedding method and wavelength of these photo-stable nanocrystals depends on their
the post-embedding method. For both methods, correlative core size, it is possible to label samples with fluorescently
approaches have been reported (Watanabe et al., 2011; Kopek different QDs, and then to distinguish individual QDs in electron
et al., 2012). In the pre-embedding method, the antigen–antibody micrographs by size (Giepmans et al., 2005; Kuipers et al., 2015).
reaction is performed before (i.e., pre) plastic embedding and A minor disadvantage of QDs, however, lies in the difficulty to
subsequent ultrathin sectioning. While better for preserving precisely measure sizes of small QDs due to its low contrast
ultra-structures, this approach is often hindered by poor in EM (Brown and Verkade, 2010). While widely used, both
FIGURE 1 | Examples of in situ and in vitro correlative light electron microscopy (CLEM) approaches. (A) In situ CLEM of confocal and transmission
electron microscopy (TEM) images showing axonal retreat from neuromuscular junctions. Top left image shows confocal image depicting axonal bulb (arrow) present
25 um from the neuromuscular junction site (in red). Middle left image illustrating surface rendering of the bulb indicated above. Top right image shows TEM, shows
that the bulb is sheathed by Schwann cells. TEM images at the bottom depict neurofilament disorganization in axonal bulb. (B) In situ CLEM of confocal and
scanning electron microscopy (SEM) images of 70 nm section from the mouse cerebral cortex. From left to right: immunostaining of ultrathin sections for tubulin,
GABA, SNAP-25, β-actin, and SEM image. Below, the boxed region is shown at a higher magnification. (C) In vitro CLEM of confocal and TEM images of
hippocampal neurons. Ultrastructure of hα-Syn inclusions in cultured neurons from SNCA+/− mice shown in TEM (top), confocal image (middle), and as merged
CLEM image (bottom). For illustration, the nucleus is rendered in blue, the cytosol in purple, and inclusions in yellow. To the right, high-resolution images of inclusions
(red boxes 1–3), with arrowheads indicating filamentous structures, are shown. (D) In vitro CLEM of confocal and SEM images of cultured hippocampal neurons.
Alignment of SEM and fluorescence signal for actin in cultured hippocampal neurons (top) and magnified sections of actin-rich convoluted nodes that form along
dendritic arbors (bottom) are shown. Scale bars: (A) 25 µm at the top left, 1 µm at the top right, 0.25 µm at the bottom; (B) 10 µm at the top and 2 µm at the
bottom; (C) 5 µm; (D) 2 µm. Pictures reprinted with permission from: (A) (Bishop et al., 2004); (B) (Micheva and Smith, 2007); (C) (Fares et al., 2016); (D) (Galic
et al., 2014).
FIGURE 2 | Examples of markers commonly used for CLEM. (A) To the left, cartoon depicting immune-gold labeling of genetically encoded fluorescent protein.
Note difference between the relative position of signal in fluorescence (yellow) and electron microscope (blue) images caused by antibodies. To the right, an example
showing fluorescence and electron micrographs of HIV particle labeled with MA-EGFP on MDCK cells expressing RFP-tagged Histone 2B. (B) To the left, cartoon
depicting protein and signal from QD in fluorescence (yellow) and electron microscope (blue) images. Note difference between protein epitope recognized by
antibody and QD signal position. To the right, example depicting RFL6 fibroblasts fixed and stained with primary antibodies followed by secondary antibodies linked
to QDs. QDs identify Cx43 at gap junctions and trafficking intermediates (green) and α-tubulin in microtubules (red). (C) To the left, cartoon depicting genetically
encoded MiniSOG, as well as the relative position and signal shape for fluorescence (yellow) and electron microscope (blue) images. To the right, an example
showing fluorescence and electron micrographs of HeLa cells expressing miniSOG labeled α-actinin. (D) To the left, cartoon depicting HaloTag labeling of protein as
well as position and signal shape for fluorescence (yellow) and electron microscope (blue) images. To the right, an example showing fluorescence and electron
micrographs of Hela cell transfected with Palmitoyl-HaloTag-meGFP. (E) Dendrites of medium-size spiny neurons in the rat neostriatum labeled with
membrane-targeted GFP and immunolabeled with Cy5 against vesicular glutamate transporter2 (VGluT2; top). After detection by fluorescence microscopy, GFP and
VGluT2 immunoreactivities were further developed for focused ion-beam SEM (FIB-SEM) via immunogold/silver enhancement and immunoperoxidase/DAB
methods, respectively (bottom). Scale bars: (A) 5 µm to the left and 100 nm to the right; (B) 5 µm; (C) 2 µm to the left and 1 µm to the right; (D) 1 µm to the left
and 500 nm to the right; (E) 3 µm. Pictures reprinted with permission from: (A) (Kukulski et al., 2011); (B) (Giepmans et al., 2005); (C) (Shu et al., 2011); (D) (Liss
et al., 2015); (E) (Sonomura et al., 2013).
methods are hampered by a localization error due to spatial post-embedding strategies. In both approaches, the region of
separation between the protein epitope and the electro-dense interest in the embedded sample is identified using a previously
particle that is visualized in light and electron micrographs determined set of reference points. Here both, biological
(Kandela et al., 2007; van Donselaar et al., 2007). To bypass (e.g., blood vessels and dirt) and artificial (e.g., silica beads
this imprecision, Aptamers directed against the fluorescence and gold markers) landmarks have proven to be useful for
tag (Shui et al., 2012) linked to gold (Javier et al., 2008; navigation. Once identified, individual EM sections are made
Chang et al., 2013), or nano-bodies directed against fluorescent and, if desired, 3D reconstruction of individual slices can be
proteins that can be conjugated to gold nanoparticles (Van de prepared by manual or automatic alignment using fiduciary
Broek et al., 2011; Ries et al., 2012) may provide promising landmarks present in adjacent sections (Bishop et al., 2004;
alternatives. Bock et al., 2011; Lee et al., 2016). Finally, electron micrographs
Orthogonal to these approaches, several strategies have (2D or 3D) are merged with the corresponding fluorescence
emerged that take advantage of singlet oxygen generators. images either manually or automatically using large features
Genetically encoded singlet oxygen generators are of particular visible in both images (e.g., blood vessels and silica beads) for
relevance for samples that are investigated ‘en block’ (i.e., coarse alignment. For subsequent fine-adjustment, subcellular
SBEM and FIB-SEM) and thus preclude post-staining of structures (e.g., gold fiduciary markers, nucleus, and filopodia)
individual sections. One such example is the genetically can be used. To estimate how accurate the alignment is, cross-
encoded miniSOG (Figure 2C), which not only fluoresces correlation analysis between light and electron micrographs can
when illuminated by blue light, but also yields an osmophilic be performed.
reaction product by catalyzing the polymerization of 3,30 - In vitro CLEM approaches permit the generation of additional
Diaminobenzidine (DAB) into electron-dense polymers that are reference systems for image alignment on the plating substrate
detectable in electron micrographs (Shu et al., 2011). The same (i.e., glass and plastic). For example, a unique reference point
strategy is also applicable using the Halo-Tag system, where a can be created at the position where the fluorescence image
modified haloalkene dehalogenase (Los et al., 2008) is used to is acquired, using for instance laser etching (Colombelli et al.,
covalently bind to specific ligands such as tetramethylrhodamine 2008; Bishop et al., 2011; Urwyler et al., 2015) or a scratching
(TMR; Figure 2D). Like miniSOG, TMR fluoresces upon device such as diamond objective markers (Sochacki et al.,
illumination and is visible in EM due to DAB oxidation 2014). Upon processing, this reference point can then be used
(Liss et al., 2015). Other self-labeling systems are provided for navigation on the electron microscope and for subsequent
by SNAP/CLIP-tags, in which the enzyme domain is bound image alignment. Another popular reference system relies on
to the protein of interest and labeled with a cell–permeable pre-formed structured (Svitkina, 2009; Benedetti et al., 2014)
fluorescent ligand (Liss et al., 2015). Finally, cells can also and stochastic (Begemann et al., 2015) micro-patterns or fiducial
be transfected with APEX, a non-fluorescent peroxidase that landmarks (Kukulski et al., 2011) for re-finding the region of
withstands strong fixation to yield light-independently EM interest on the EM as well as subsequent image alignment.
contrast (Martell et al., 2012). However, considering that the Intriguingly, such reference points are not only suitable for
reaction product can diffuse, these approaches may yield best manual alignment, but also for software-assisted solutions
results when staining enclosed structures. Notably, multiple (Begemann et al., 2015), rendering these approaches an attractive
procedures can be combined within a sample for visualization entry point compared to more sophisticated CLEM strategies.
(Figure 2E). Finally, simultaneous fluorescence and electron imaging may
provide an alternative strategy. Here, a number of custom-made
(Liv et al., 2013; Nishiyama et al., 2014; Peddie et al., 2014; de
Strategies of Aligning Fluorescence Boer et al., 2015) and commercial (e.g., from Fei and Zeiss)
Image and Electron Micrograph in CLEM instruments have been presented. These instruments, which are
Correlation of fluorescence images and electron micrographs equally suitable for in situ and in vitro samples, integrate light-
can be challenging for a number of reasons. For one, the and electron imaging in one apparatus (iCLEM), thus ensuring
region of interest can cover less than 0.0004% of the total dual imaging of the region of interest without need of later
sample area (Begemann et al., 2015), thus offering only image alignment. However, as sample preparation for iCLEM
on a limited number of reference points for navigation. is restricted to techniques suitable for parallel LM and EM
And even after the region of interest acquired at the light (Agronskaia et al., 2008), and due to the limited availability
microscope is found again on the electron microscope, rotation, of such instruments, iCLEM has not yet reached the optional
magnification, and tilting angle still need to be adjusted for image users.
alignment. This is of particular relevance in pre-embedding
approaches, where, unlike to post-embedding strategies, the
angle of the respective acquisition planes may differ. To CLEM–POTENTIAL AND LIMITATIONS
tackle these obstacles, several strategies for navigation and
image alignment have emerged for in situ and in vitro In this review, we have discussed possible combinations for
approaches. correlating optical and electron microscopy approaches. We
For in situ CLEM, various fiduciary markers have been mentioned how improvements in imaging techniques, sample
described that are equally suitable for pre-embedding and preparation, markers, and image alignment have facilitated
FIGURE 3 | Graphical summary of existing CLEM approaches. Electron microscopy (blue) and light-based techniques (yellow) that were successfully combined
in correlative approaches are depicted by white lines connecting the respective methods. References corresponding to individual publications using specific CLEM
approaches (numbered 1–50) can be found in the Supplementary Materials (Supplementary Table S1).
development of novel CLEM approaches, and showed examples continuous increase in sophistication and computational power
where these strategies were successfully used for structure- used for image analysis have captured the imagination of people.
function analysis in neurons and synapses. In order to visualize While it remains elusive whether these and other emerging
the combinatorial potential, we provide a chart summarizing techniques will merge or replace CLEM-based studies, they
published CLEM pairings (Figure 3). are exemplary for an ongoing push toward visualization and
While correlative approaches have revealed aspects of comprehensive analysis of biological function on the structural
neuronal and synaptic function that would not have been level.
possible without it, challenges still remain due to the limited Taken together, the possibilities of applying correlative
temporal and spatial resolution of CLEM. From this perspective, approaches to study neuronal and synapse function are growing
recent work on liquid cell electron microscopy (Ross, 2015), thanks to the continuous progress in techniques, tools and
where electron-permeable membranes are used to host aqueous protocols for combining these two types of microscopy.
solutions under atmospheric pressure conditions for TEM and Alas, while correlative approaches in neurobiology have
SEM imaging (Danilatos et al., 2011; Ominami et al., 2015), substantially increased in quality and availability, truly artifact-
cryo-ET to study the supramolecular architecture of cellular free preparation techniques for precise molecule-localization are
structures (Medalia et al., 2002; Beck et al., 2007), and the yet to be achieved.
REFERENCES Bosch, C., Martinez, A., Masachs, N., Teixeira, C. M., Fernaud, I., Ulloa, F.,
et al. (2015). FIB/SEM technology and high-throughput 3D reconstruction of
Abbe (1883). The relation of aperture and power in the microscope. J. R. Microsc. dendritic spines and synapses in GFP-labeled adult-generated neurons. Front.
Soc. 3, 790–812. doi: 10.1111/j.1365-2818.1883.tb05956.x Neuroanat. 9:60. doi: 10.3389/fnana.2015.00060
Adams, S. R., Harootunian, A. T., Buechler, Y. J., Taylor, S. S., and Tsien, R. Y. Boyden, E. S., Zhang, F., Bamberg, E., Nagel, G., and Deisseroth, K. (2005).
(1991). Fluorescence ratio imaging of cyclic AMP in single cells. Nature 349, Millisecond-timescale, genetically targeted optical control of neural activity.
694–697. doi: 10.1038/349694a0 Nat. Neurosci. 8, 1263–1268. doi: 10.1038/nn1525
Agronskaia, A. V., Valentijn, J. A., van Driel, L. F., Schneijdenberg, C. T., Humbel, Briggman, K. L., and Bock, D. D. (2012). Volume electron microscopy for
B. M., van Bergen en Henegouwen, P. M., et al. (2008). Integrated fluorescence neuronal circuit reconstruction. Curr. Opin. Neurobiol. 22, 154–161. doi:
and transmission electron microscopy. J. Struct. Biol. 164, 183–189. doi: 10.1016/j.conb.2011.10.022
10.1016/j.jsb.2008.07.003 Briggman, K. L., Helmstaedter, M., and Denk, W. (2011). Wiring specificity
Akagi, T., Ishida, K., Hanasaka, T., Hayashi, S., Watanabe, M., Hashikawa, T., in the direction-selectivity circuit of the retina. Nature 471, 183–188. doi:
et al. (2006). Improved methods for ultracryotomy of CNS tissue for 10.1038/nature09818
ultrastructural and immunogold analyses. J. Neurosci. Methods 153, 276–282. Brown, E., and Verkade, P. (2010). The use of markers for correlative light electron
doi: 10.1016/j.jneumeth.2005.11.007 microscopy. Protoplasma 244, 91–97. doi: 10.1007/s00709-010-0165-1
Al Jord, A., Lemaitre, A. I., Delgehyr, N., Faucourt, M., Spassky, N., and Meunier, A. Cao, C., Lu, S., Kivlin, R., Wallin, B., Card, E., Bagdasarian, A., et al.
(2014). Centriole amplification by mother and daughter centrioles differs in (2008). AMP-activated protein kinase contributes to UV- and H2O2-induced
multiciliated cells. Nature 516, 104–107. doi: 10.1038/nature13770 apoptosis in human skin keratinocytes. J. Biol. Chem. 283, 28897–28908. doi:
Alivisatos, A. P. (1996). Semiconductor clusters, nanocrystals, and quantum dots. 10.1074/jbc.M804144200
Science 271, 933–937. doi: 10.1126/science.271.5251.933 Castejon, O. J. (1996). Contribution of conventional and high resolution scanning
Ashkin, A. (1970). Acceleration and trapping of particles by radiation pressure. electron microscopy and cryofracture technique to the study of cerebellar
Phys. Rev. Lett. 24, 156–159. doi: 10.1103/PhysRevLett.24.156 synaptic junctions. Scanning Microsc. 10, 177–186.
Axelrod, D. (1981). Cell-substrate contacts illuminated by total internal reflection Castejon, O. J., and Caraballo, A. J. (1980). Application of cryofracture and SEM to
fluorescence. J. Cell Biol. 89, 141–145. doi: 10.1083/jcb.89.1.141 the study of human cerebellar cortex. Scan. Electron Microsc. 4, 197–207.
Banker, G. A., and Cowan, W. M. (1977). Rat hippocampal neurons in dispersed Chalfie, M., Tu, Y., Euskirchen, G., Ward, W. W., and Prasher, D. C. (1994). Green
cell culture. Brain Res. 126, 297–342. doi: 10.1016/0006-8993(77)90594-7 fluorescent protein as a marker for gene expression. Science 263, 802–805. doi:
Beck, M., Lucic, V., Forster, F., Baumeister, W., and Medalia, O. (2007). Snapshots 10.1126/science.8303295
of nuclear pore complexes in action captured by cryo-electron tomography. Chang, Y. C., Yang, C. Y., Sun, R. L., Cheng, Y. F., Kao, W. C., and Yang,
Nature 449, 611–615. doi: 10.1038/nature06170 P. C. (2013). Rapid single cell detection of Staphylococcus aureus by aptamer-
Begemann, I., Viplav, A., Rasch, C., and Galic, M. (2015). Stochastic micro-pattern conjugated gold nanoparticles. Sci. Rep. 3, 1863. doi: 10.1038/srep01863
for automated correlative fluorescence – scanning electron microscopy. Sci. Collman, F., Buchanan, J., Phend, K. D., Micheva, K. D., Weinberg, R. J., and Smith,
Rep. 5, 17973. doi: 10.1038/srep17973 S. J. (2015). Mapping synapses by conjugate light-electron array tomography.
Benedetti, L., Sogne, E., Rodighiero, S., Marchesi, D., Milani, P., and Francolini, M. J. Neurosci. 35, 5792–5807. doi: 10.1523/JNEUROSCI.4274-14.2015
(2014). Customized patterned substrates for highly versatile correlative light- Colombelli, J., Tangemo, C., Haselman, U., Antony, C., Stelzer, E. H.,
scanning electron microscopy. Sci. Rep. 4, 7033. doi: 10.1038/srep07033 Pepperkok, R., et al. (2008). A correlative light and electron microscopy method
Betzig, E., and Chichester, R. J. (1993). Single molecules observed by based on laser micropatterning and etching. Methods Mol. Biol. 457, 203–213.
near-field scanning optical microscopy. Science 262, 1422–1425. doi: doi: 10.1007/978-1-59745-261-8_15
10.1126/science.262.5138.1422 Combs, C. A. (2010). Fluorescence microscopy: a concise guide to current
Betzig, E., Patterson, G. H., Sougrat, R., Lindwasser, O. W., Olenych, S., Bonifacino, imaging methods. Curr. Protoc. Neurosci. Chapter 2, Unit2.1. doi:
J. S., et al. (2006). Imaging intracellular fluorescent proteins at nanometer 10.1002/0471142301.ns0201s50
resolution. Science 313, 1642–1645. doi: 10.1126/science.1127344 Crewe, A. V., Wall, J., and Langmore, J. (1970). Visibility of single atoms. Science
Bishop, D., Nikic, I., Brinkoetter, M., Knecht, S., Potz, S., Kerschensteiner, M., 168, 1338–1340. doi: 10.1126/science.168.3937.1338
et al. (2011). Near-infrared branding efficiently correlates light and electron Crowther, R. A., Derosier, D. J., and Klug, A. (1970). Reconstruction of
microscopy. Nat. Methods 8, 568–570. doi: 10.1038/nmeth.1622 3 dimensional structure from projections and its application to electron
Bishop, D. L., Misgeld, T., Walsh, M. K., Gan, W. B., and Lichtman, J. W. (2004). microscopy. Proc. R. Soc. Lond. A 317, 319–340. doi: 10.1098/rspa.1970.0119
Axon branch removal at developing synapses by axosome shedding. Neuron 44, Dahl, R., and Staehelin, L. A. (1989). High-pressure freezing for the preservation of
651–661. doi: 10.1016/j.neuron.2004.10.026 biological structure – theory and practice. J. Electron Microsc. Tech. 13, 165–174.
Blazquez-Llorca, L., Hummel, E., Zimmerman, H., Zou, C., Burgold, S., Rietdorf, J., doi: 10.1002/jemt.1060130305
et al. (2015). Correlation of two-photon in vivo imaging and FIB/SEM Danilatos, G., Rattenberger, J., and Dracopoulos, V. (2011). Beam transfer
microscopy. J. Microsc. 259, 129–136. doi: 10.1111/jmi.12231 characteristics of a commercial environmental SEM and a low vacuum SEM.
Bock, D. D., Lee, W. C., Kerlin, A. M., Andermann, M. L., Hood, G., Wetzel, J. Microsc. 242, 166–180. doi: 10.1111/j.1365-2818.2010.03455.x
A. W., et al. (2011). Network anatomy and in vivo physiology of visual cortical Davila, H. V., Salzberg, B. M., Cohen, L. B., and Waggoner, A. S. (1973).
neurons. Nature 471, 177–182. doi: 10.1038/nature09802 A large change in axon fluorescence that provides a promising method
for measuring membrane potential. Nat. New Biol. 241, 159–160. doi: Harris, K. M., Perry, E., Bourne, J., Feinberg, M., Ostroff, L., and Hurlburt, J. (2006).
10.1038/newbio241159a0 Uniform serial sectioning for transmission electron microscopy. J. Neurosci. 26,
de Boer, P., Hoogenboom, J. P., and Giepmans, B. N. (2015). Correlated light and 12101–12103. doi: 10.1523/JNEUROSCI.3994-06.2006
electron microscopy: ultrastructure lights up! Nat. Methods 12, 503–513. doi: Hayworth, K. J., Kasthuri, N., Schalek, R., and Lichtman, J. (2006). Automating
10.1038/nmeth.3400 the collection of ultrathin serial sections for large volume TEM reconstructions.
De Rosier, D. J., and Klug, A. (1968). Reconstruction of three dimensional Microsc. Microanal. 12, 86–87. doi: 10.1017/S1431927606066268
structures from electron micrographs. Nature 217, 130–134. doi: Hayworth, K. J., Morgan, J. L., Schalek, R., Berger, D. R., Hildebrand, D. G.,
10.1038/217130a0 and Lichtman, J. W. (2014). Imaging ATUM ultrathin section libraries with
Deerinck, T. J., Martone, M. E., Lev-Ram, V., Green, D. P., Tsien, R. Y., Spector, WaferMapper: a multi-scale approach to EM reconstruction of neural circuits.
D. L., et al. (1994). Fluorescence photooxidation with eosin: a method for high Front. Neural Circuits 8:68. doi: 10.3389/fncir.2014.00068
resolution immunolocalization and in situ hybridization detection for light and Heim, R., Prasher, D. C., and Tsien, R. Y. (1994). Wavelength mutations and
electron microscopy. J. Cell Biol. 126, 901–910. doi: 10.1083/jcb.126.4.901 posttranslational autoxidation of green fluorescent protein. Proc. Natl. Acad.
DeKosky, S. T., and Scheff, S. W. (1990). Synapse loss in frontal cortex biopsies Sci. U.S.A. 91, 12501–12504. doi: 10.1073/pnas.91.26.12501
in Alzheimer’s disease: correlation with cognitive severity. Ann. Neurol. 27, Hell, S. W., and Wichmann, J. (1994). Breaking the diffraction resolution limit by
457–464. doi: 10.1002/ana.410270502 stimulated emission: stimulated-emission-depletion fluorescence microscopy.
Denk, W., and Horstmann, H. (2004). Serial block-face scanning electron Opt. Lett. 19, 780–782. doi: 10.1364/OL.19.000780
microscopy to reconstruct three-dimensional tissue nanostructure. PLoS Biol. Heuser, J. E., and Reese, T. S. (1973). Evidence for recycling of synaptic vesicle
2:e329. doi: 10.1371/journal.pbio.0020329 membrane during transmitter release at the frog neuromuscular junction. J. Cell
Denk, W., Strickler, J. H., and Webb, W. W. (1990). Two-photon laser scanning Biol. 57, 315–344. doi: 10.1083/jcb.57.2.315
fluorescence microscopy. Science 248, 73–76. doi: 10.1126/science.2321027 Heuser, J. E., Reese, T. S., and Landis, D. M. (1976). Preservation of synaptic
Derobertis, E. D. P., and Bennett, H. S. (1955). Some features of the submicroscopic structure by rapid freezing. Cold Spring Harb. Symp. Quant. Biol. 40, 17–24.
morphology of synapses in frog and earthworm. J. Biophys. Biochem. Cytol. 1, doi: 10.1101/SQB.1976.040.01.004
47–58. doi: 10.1083/jcb.1.1.47 Heuser, J. E., and Salpeter, S. R. (1979). Organization of acetylcholine receptors
Desaki, J., and Uehara, Y. (1981). The overall morphology of neuromuscular in quick-frozen, deep-etched, and rotary-replicated Torpedo postsynaptic
junctions as revealed by scanning electron microscopy. J. Neurocytol. 10, membrane. J. Cell Biol. 82, 150–173. doi: 10.1083/jcb.82.1.150
101–110. doi: 10.1007/BF01181747 Hollander, H. (1970). The section embedding (SE) technique. A new method
Ellis-Davies, G. C. (2011). A practical guide to the synthesis of dinitroindolinyl- for the combined light microscopic and electron microscopic examination of
caged neurotransmitters. Nat. Protoc. 6, 314–326. doi: 10.1038/nprot.2010.193 central nervous tissue. Brain Res. 20, 39–47.
Ellisdavies, G. C. R., and Kaplan, J. H. (1994). Nitrophenyl-Egta, a photolabile Hoppe, W., Gassmann, J., Hunsmann, N., Schramm, H. J., and Sturm, M. (1974).
chelator that selectively binds Ca2+ with high-affinity and releases it 3-dimensional reconstruction of individual negatively stained yeast fatty-acid
rapidly upon photolysis. Proc. Natl. Acad. Sci. U.S.A. 91, 187–191. doi: synthetase molecules from tilt series in electron-microscope. Hoppe Seylers Z.
10.1073/pnas.91.1.187 Physiol. Chem. 355, 1483–1487.
Engel, A. (1978). The STEM: an attractive tool for the biologist. Ultramicroscopy 3, Huang, B., Babcock, H., and Zhuang, X. (2010). Breaking the diffraction
355–357. doi: 10.1016/S0304-3991(78)80052-7 barrier: super-resolution imaging of cells. Cell 143, 1047–1058. doi:
Engelbrecht, C. J., and Stelzer, E. H. (2006). Resolution enhancement in 10.1016/j.cell.2010.12.002
a light-sheet-based microscope (SPIM). Opt. Lett. 31, 1477–1479. doi: Hunziker, E. B., Herrmann, W., Schenk, R. K., Mueller, M., and Moor, H. (1984).
10.1364/OL.31.001477 Cartilage ultrastructure after high pressure freezing, freeze substitution, and
Ercius, P., Alaidi, O., Rames, M. J., and Ren, G. (2015). Electron tomography: a low temperature embedding. I. Chondrocyte ultrastructure–implications for
three-dimensional analytic tool for hard and soft materials research. Adv. Mater. the theories of mineralization and vascular invasion. J. Cell Biol. 98, 267–276.
27, 5638–5663. doi: 10.1002/adma.201501015 Javier, D. J., Nitin, N., Levy, M., Ellington, A., and Richards-Kortum, R. (2008).
Erni, R., Rossell, M. D., Kisielowski, C., and Dahmen, U. (2009). Atomic-resolution Aptamer-targeted gold nanoparticles as molecular-specific contrast agents for
imaging with a sub-50-pm electron probe. Phys. Rev. Lett. 102, 096101. doi: reflectance imaging. Bioconjug. Chem. 19, 1309–1312. doi: 10.1021/bc8001248
10.1103/PhysRevLett.102.096101 Kandela, I. K., and Albrecht, R. M. (2007). Fluorescence quenching by colloidal
Fares, M. B., Maco, B., Oueslati, A., Rockenstein, E., Ninkina, N., Buchman, V. L., heavy metals nanoparticles: implications for correlative fluorescence and
et al. (2016). Induction of de novo alpha-synuclein fibrillization in a neuronal electron microscopy studies. Scanning 29, 152–161. doi: 10.1002/sca.20055
model for Parkinson’s disease. Proc. Natl. Acad. Sci. U.S.A. 113, E912–E921. doi: Kandela, I. K., Bleher, R., and Albrecht, R. M. (2007). Multiple correlative
10.1073/pnas.1512876113 immunolabeling for light and electron microscopy using fluorophores
Fenno, L., Yizhar, O., and Deisseroth, K. (2011). The development and application and colloidal metal particles. J. Histochem. Cytochem. 55, 983–990. doi:
of optogenetics. Annu. Rev. Neurosci. 34, 389–412. doi: 10.1146/annurev-neuro- 10.1369/jhc.6A7124.2007
061010-113817 Kasthuri, N., Hayworth, K., Lichtman, J., Erdman, N., and Ackerley, C. A.
Fifkova, E., and Delay, R. J. (1982). Cytoplasmic actin in neuronal processes (2007). New technique for ultra-thin serial brain section imaging using
as a possible mediator of synaptic plasticity. J. Cell Biol. 95, 345–350. doi: scanning electron microscopy. Microsc. Microanal. 13, 26–27. doi:
10.1083/jcb.95.1.345 10.1017/S1431927607078002
Galic, M., Jeong, S., Tsai, F. C., Joubert, L. M., Wu, Y. I., Hahn, K. M., et al. Kasthuri, N., Hayworth, K. J., Berger, D. R., Schalek, R. L., Conchello, J. A.,
(2012). External push and internal pull forces recruit curvature-sensing N-BAR Knowles-Barley, S., et al. (2015). Saturated reconstruction of a volume of
domain proteins to the plasma membrane. Nat. Cell Biol. 14, 874–881. doi: neocortex. Cell 162, 648–661. doi: 10.1016/j.cell.2015.06.054
10.1038/ncb2533 Keller, P. J., Schmidt, A. D., Wittbrodt, J., and Stelzer, E. H. (2008). Reconstruction
Galic, M., Tsai, F. C., Collins, S. R., Matis, M., Bandara, S., and Meyer, T. of zebrafish early embryonic development by scanned light sheet microscopy.
(2014). Dynamic recruitment of the curvature-sensitive protein ArhGAP44 to Science 322, 1065–1069. doi: 10.1126/science.1162493
nanoscale membrane deformations limits exploratory filopodia initiation in Kilgore, J. A., Dolman, N. J., and Davidson, M. W. (2013). A review of reagents
neurons. Elife 3, e03116. doi: 10.7554/eLife.03116 for fluorescence microscopy of cellular compartments and structures, Part II:
Giepmans, B. N., Deerinck, T. J., Smarr, B. L., Jones, Y. Z., and Ellisman, reagents for non-vesicular organelles. Curr. Protoc. Cytom. 66, Unit 12.31. doi:
M. H. (2005). Correlated light and electron microscopic imaging of multiple 10.1002/0471142956.cy1231s66
endogenous proteins using Quantum dots. Nat. Methods 2, 743–749. doi: Kim, D., Deerinck, T. J., Sigal, Y. M., Babcock, H. P., Ellisman, M. H.,
10.1038/nmeth791 and Zhuang, X. (2015). Correlative stochastic optical reconstruction
Gray, E. G. (1959). Axo-somatic and axo-dendritic synapses of the cerebral cortex: microscopy and electron microscopy. PLoS ONE 10:e0124581. doi:
an electron microscope study. J. Anat. 93, 420–433. 10.1371/journal.pone.0124581
Klar, T. A., Engel, E., and Hell, S. W. (2001). Breaking Abbe’s diffraction resolution visualized by cryoelectron tomography. Science 298, 1209–1213. doi:
limit in fluorescence microscopy with stimulated emission depletion beams of 10.1126/science.1076184
various shapes. Phys. Rev. E Stat. Nonlin. Soft. Matter. Phys. 64, 066613. doi: Merchan-Perez, A., Rodriguez, J. R., Alonso-Nanclares, L., Schertel, A., and
10.1103/PhysRevE.64.066613 Defelipe, J. (2009). Counting Synapses Using FIB/SEM microscopy: a true
Knott, G., and Genoud, C. (2013). Is EM dead? J. Cell Sci. 126, 4545–4552. doi: revolution for ultrastructural volume reconstruction. Front. Neuroanat. 3:18.
10.1242/jcs.124123 doi: 10.3389/neuro.05.018.2009
Knott, G., Marchman, H., Wall, D., and Lich, B. (2008). Serial section scanning Micheva, K. D., and Smith, S. J. (2007). Array tomography: a new tool for imaging
electron microscopy of adult brain tissue using focused ion beam milling. the molecular architecture and ultrastructure of neural circuits. Neuron 55,
J. Neurosci. 28(Pt. 20), 2959–2964. doi: 10.1523/JNEUROSCI.3189-07.2008 25–36. doi: 10.1016/j.neuron.2007.06.014
Kopek, B. G., Shtengel, G., Xu, C. S., Clayton, D. A., and Hess, H. F. (2012). Miesenbock, G., De Angelis, D. A., and Rothman, J. E. (1998). Visualizing secretion
Correlative 3D superresolution fluorescence and electron microscopy reveal the and synaptic transmission with pH-sensitive green fluorescent proteins. Nature
relationship of mitochondrial nucleoids to membranes. Proc. Natl. Acad. Sci. 394, 192–195. doi: 10.1038/28190
U.S.A. 109, 6136–6141. doi: 10.1073/pnas.1121558109 Minsky, M. (1988). Memoir on inventing the confocal scanning microscope.
Kuipers, J., de Boer, P., and Giepmans, B. N. (2015). Scanning EM of non-heavy Scanning 10, 128–138. doi: 10.1002/sca.4950100403
metal stained biosamples: large-field of view, high contrast and highly efficient Minta, A., and Tsien, R. Y. (1989). Fluorescent indicators for cytosolic sodium.
immunolabeling. Exp. Cell Res. 337, 202–207. doi: 10.1016/j.yexcr.2015.07.012 J. Biol. Chem. 264, 19449–19457.
Kukulski, W., Schorb, M., Welsch, S., Picco, A., Kaksonen, M., and Briggs, Miranda, K., Girard-Dias, W., Attias, M., de Souza, W., and Ramos, I. (2015).
J. A. (2011). Correlated fluorescence and 3D electron microscopy with Three dimensional reconstruction by electron microscopy in the life sciences:
high sensitivity and spatial precision. J. Cell Biol. 192, 111–119. doi: an introduction for cell and tissue biologists. Mol. Reprod. Dev. 82, 530–547.
10.1083/jcb.201009037 doi: 10.1002/mrd.22455
Kuwajima, M., Mendenhall, J. M., Lindsey, L. F., and Harris, K. M. (2013). Mishchenko, Y., Hu, T., Spacek, J., Mendenhall, J., Harris, K. M., and
Automated transmission-mode scanning electron microscopy (tSEM) for Chklovskii, D. B. (2010). Ultrastructural analysis of hippocampal
large volume analysis at nanoscale resolution. PLoS ONE 8:e59573. doi: neuropil from the connectomics perspective. Neuron 67, 1009–1020. doi:
10.1371/journal.pone.0059573 10.1016/j.neuron.2010.08.014
Landis, D. M. D., and Reese, T. S. (1983). Cytoplasmic organization in cerebellar Morgan, J. L., Berger, D. R., Wetzel, A. W., and Lichtman, J. W. (2016). The fuzzy
dendritic spines. J. Cell Biol. 97, 1169–1178. doi: 10.1083/jcb.97.4.1169 logic of network connectivity in mouse visual Thalamus. Cell 165, 192–206. doi:
Lang, T. (2003). Imaging SNAREs at work in ‘unroofed’ cells–approaches that may 10.1016/j.cell.2016.02.033
be of general interest for functional studies on membrane proteins. Biochem. Muller-Reichert, T., and Verkade, P. (2014). Preface. Correlative light and electron
Soc. Trans. 31, 861–864. doi: 10.1042/bst0310861 microscopy II. Methods Cell Biol. 124, xvii–xviii. doi: 10.1016/B978-0-12-
Lee, W. C., Bonin, V., Reed, M., Graham, B. J., Hood, G., Glattfelder, K., et al. 801075-4.09983-3
(2016). Anatomy and function of an excitatory network in the visual cortex. Nakai, J., Ohkura, M., and Imoto, K. (2001). A high signal-to-noise Ca(2+) probe
Nature 532, 370–374. doi: 10.1038/nature17192 composed of a single green fluorescent protein. Nat. Biotechnol. 19, 137–141.
Lehmann, T., Hess, M., Wanner, G., and Melzer, R. R. (2014). Dissecting a neuron doi: 10.1038/84397
network: FIB-SEM-based 3D-reconstruction of the visual neuropils in the Nakai, Y., and Iwashita, T. (1976). Correlative light and electron microscopy of the
sea spider Achelia langi (Dohrn, 1881) (Pycnogonida). BMC Biol. 12:59. doi: frog adrenal gland cells using adjacent epon-embedded sections. Arch. Histol.
10.1186/s12915-014-0059-3 Jpn. 39, 183–191. doi: 10.1679/aohc1950.39.183
Leighton, S. B. (1981). SEM images of block faces, cut by a miniature microtome Navlakha, S., Suhan, J., Barth, A. L., and Bar-Joseph, Z. (2013). A high-throughput
within the SEM – a technical note. Scan. Electron Microsc. 233, 73–76. framework to detect synapses in electron microscopy images. Bioinformatics 29,
Lenzi, D., Runyeon, J. W., Crum, J., Ellisman, M. H., and Roberts, W. M. (1999). i9–i17. doi: 10.1093/bioinformatics/btt222
Synaptic vesicle populations in saccular hair cells reconstructed by electron Neil, M. A., Juskaitis, R., and Wilson, T. (1997). Method of obtaining optical
tomography. J. Neurosci. 19, 119–132. sectioning by using structured light in a conventional microscope. Opt. Lett.
Lichtman, J. W., and Conchello, J. A. (2005). Fluorescence microscopy. Nat. 22, 1905–1907. doi: 10.1364/OL.22.001905
Methods 2, 910–919. doi: 10.1038/nmeth817 Nishiyama, H., Koizumi, M., Ogawa, K., Kitamura, S., Konyuba, Y., Watanabe, Y.,
Liss, V., Barlag, B., Nietschke, M., and Hensel, M. (2015). Self-labelling enzymes et al. (2014). Atmospheric scanning electron microscope system with an open
as universal tags for fluorescence microscopy, super-resolution microscopy and sample chamber: configuration and applications. Ultramicroscopy 147, 86–97.
electron microscopy. Sci. Rep. 5, 17740. doi: 10.1038/srep17740 doi: 10.1016/j.ultramic.2014.06.001
Liv, N., Zonnevylle, A. C., Narvaez, A. C., Effting, A. P., Voorneveld, Ominami, Y., Kawanishi, S., Ushiki, T., and Ito, S. (2015). A novel approach to
P. W., Lucas, M. S., et al. (2013). Simultaneous correlative scanning scanning electron microscopy at ambient atmospheric pressure. Microscopy 64,
electron and high-NA fluorescence microscopy. PLoS ONE 8:e55707. doi: 97–104. doi: 10.1093/jmicro/dfu107
10.1371/journal.pone.0055707 Osamura, R. Y., Itoh, Y., and Matsuno, A. (2000). Applications of plastic
Livet, J., Weissman, T. A., Kang, H., Draft, R. W., Lu, J., Bennis, R. A., et al. (2007). embedding to electron microscopic immunocytochemistry and in situ
Transgenic strategies for combinatorial expression of fluorescent proteins in the hybridization in observations of production and secretion of peptide hormones.
nervous system. Nature 450, 56–62. doi: 10.1038/nature06293 J. Histochem. Cytochem. 48, 885–891. doi: 10.1177/002215540004800701
Los, G. V., Encell, L. P., McDougall, M. G., Hartzell, D. D., Karassina, N., Paez-Segala, M. G., Sun, M. G., Shtengel, G., Viswanathan, S., Baird, M. A.,
Zimprich, C., et al. (2008). HaloTag: a novel protein labeling technology Macklin, J. J., et al. (2015). Fixation-resistant photoactivatable fluorescent
for cell imaging and protein analysis. ACS Chem. Biol. 3, 373–382. doi: proteins for CLEM. Nat. Methods 12, 215–218. doi: 10.1038/nmeth.3225
10.1021/cb800025k Pappas, G. D., and Bennett, M. V. L. (1966). Specialized junctions involved in
Maco, B., Holtmaat, A., Cantoni, M., Kreshuk, A., Straehle, C. N., Hamprecht, electrical transmission between neurons. Ann. N. Y. Acad. Sci. 137, 495–508.
F. A., et al. (2013). Correlative in vivo 2 photon and focused ion beam doi: 10.1111/j.1749-6632.1966.tb50177.x
scanning electron microscopy of cortical neurons. PLoS ONE 8:e57405. doi: Peddie, C. J., Blight, K., Wilson, E., Melia, C., Marrison, J., Carzaniga, R.,
10.1371/journal.pone.0057405 et al. (2014). Correlative and integrated light and electron microscopy of in-
Martell, J. D., Deerinck, T. J., Sancak, Y., Poulos, T. L., Mootha, V. K., resin GFP fluorescence, used to localise diacylglycerol in mammalian cells.
Sosinsky, G. E., et al. (2012). Engineered ascorbate peroxidase as a genetically Ultramicroscopy 143, 3–14. doi: 10.1016/j.ultramic.2014.02.001
encoded reporter for electron microscopy. Nat. Biotechnol. 30, 1143–1148. doi: Penczek, P. A. (2010). Fundamentals of three-dimensional reconstruction from
10.1038/nbt.2375 projections. Methods Enzymol. 482, 1–33. doi: 10.1016/S0076-6879(10)82001-4
Medalia, O., Weber, I., Frangakis, A. S., Nicastro, D., Gerisch, G., and Perkins, G. A., Jackson, D. R., and Spirou, G. A. (2015). Resolving presynaptic
Baumeister, W. (2002). Macromolecular architecture in eukaryotic cells structure by electron tomography. Synapse 69, 268–282. doi: 10.1002/syn.21813
Phend, K. D., Rustioni, A., and Weinberg, R. J. (1995). An osmium-free method of Terasaki, M., Shemesh, T., Kasthuri, N., Klemm, R. W., Schalek, R., Hayworth,
epon embedment that preserves both ultrastructure and antigenicity for post- K. J., et al. (2013). Stacked endoplasmic reticulum sheets are connected
embedding immunocytochemistry. J. Histochem. Cytochem. 43, 283–292. doi: by helicoidal membrane motifs. Cell 154, 285–296. doi: 10.1016/j.cell.2013.
10.1177/43.3.7532656 06.031
Pierce, L., and Buseck, P. R. (1974). Electron imaging of pyrrhotite superstructures. Tischer, D., and Weiner, O. D. (2014). Illuminating cell signalling with optogenetic
Science 186, 1209–1212. doi: 10.1126/science.186.4170.1209 tools. Nat. Rev. Mol. Cell Biol. 15, 551–558. doi: 10.1038/nrm3837
Porter, K. R., Claude, A., and Fullam, E. F. (1945). A study of tissue culture cells by Tocheva, E. I., Li, Z., and Jensen, G. J. (2010). Electron cryotomography.
electron microscopy: methods and preliminary observations. J. Exp. Med. 81, Cold Spring Harb. Perspect. Biol. 2, a003442. doi: 10.1101/cshperspect.
233–246. doi: 10.1084/jem.81.3.233 a003442
Rah, J. C., Bas, E., Colonell, J., Mishchenko, Y., Karsh, B., Fetter, R. D., et al. Tsien, R. Y. (1980). New calcium indicators and buffers with high selectivity
(2013). Thalamocortical input onto layer 5 pyramidal neurons measured using against magnesium and protons: design, synthesis, and properties of prototype
quantitative large-scale array tomography. Front. Neural Circuits 7:177. doi: structures. Biochemistry 19, 2396–2404. doi: 10.1021/bi00552a018
10.3389/fncir.2013.00177 Tsien, R. Y. (1981). A non-disruptive technique for loading calcium buffers and
Rayleigh (1879). Investigations in optics, with special reference to the spectroscope. indicators into cells. Nature 290, 527–528. doi: 10.1038/290527a0
Phil. Mag. 8, 261–274. doi: 10.1080/14786447908639715 Tsien, R. Y., Pozzan, T., and Rink, T. J. (1982). Calcium homeostasis in intact
Ries, J., Kaplan, C., Platonova, E., Eghlidi, H., and Ewers, H. (2012). A simple, lymphocytes: cytoplasmic free calcium monitored with a new, intracellularly
versatile method for GFP-based super-resolution microscopy via nanobodies. trapped fluorescent indicator. J. Cell Biol. 94, 325–334. doi: 10.1083/jcb.94.2.325
Nat. Methods 9, 582–584. doi: 10.1038/nmeth.1991 Urwyler, O., Izadifar, A., Dascenco, D., Petrovic, M., He, H. H., Ayaz, D.,
Ross, F. M. (2015). Opportunities and challenges in liquid cell electron microscopy. et al. (2015). Investigating CNS synaptogenesis at single-synapse resolution
Science 350, aaa9886. doi: 10.1126/science.aaa9886 by combining reverse genetics with correlative light and electron microscopy.
Rost, B. R., Schneider, F., Grauel, M. K., Wozny, C., Bentz, C. G., Blessing, A., Development 142, 394–405. doi: 10.1242/dev.115071
et al. (2015). Optogenetic acidification of synaptic vesicles and lysosomes. Nat. Van de Broek, B., Devoogdt, N., D’Hollander, A., Gijs, H. L., Jans, K., Lagae, L.,
Neurosci. 18, 1845–1852. doi: 10.1038/nn.4161 et al. (2011). Specific cell targeting with nanobody conjugated branched
Ruska, E. (1987). Nobel lecture. The development of the electron microscope and gold nanoparticles for photothermal therapy. ACS Nano 5, 4319–4328. doi:
of electron microscopy. Biosci. Rep. 7, 607–629. 10.1021/nn1023363
Rust, M. J., Bates, M., and Zhuang, X. W. (2006). Sub-diffraction-limit imaging van Donselaar, E., Posthuma, G., Zeuschner, D., Humbel, B. M., and Slot, J. W.
by stochastic optical reconstruction microscopy (STORM). Nat. Methods 3, (2007). Immunogold labeling of cryosections from high-pressure frozen cells.
793–795. doi: 10.1038/nmeth929 Traffic 8, 471–485. doi: 10.1111/j.1600-0854.2007.00552.x
Shimomura, O., Johnson, F. H., and Saiga, Y. (1962). Extraction, purification Vernon-Parry, K. D. (2000). Scanning electron microscopy: an Introduction. III-Vs
and properties of aequorin, a bioluminescent protein from the luminous Rev. 13, 40–44. doi: 10.1016/S0961-1290(00)80006-X
hydromedusan, Aequorea. J. Cell Comp. Physiol. 59, 223–239. doi: Wang, S. S. H., and Augustine, G. J. (1995). Confocal imaging and local photolysis
10.1002/jcp.1030590302 of caged compounds – dual probes of synaptic function. Neuron 15, 755–760.
Shu, X., Lev-Ram, V., Deerinck, T. J., Qi, Y., Ramko, E. B., Davidson, M. W., doi: 10.1016/0896-6273(95)90167-1
et al. (2011). A genetically encoded tag for correlated light and electron Watanabe, S., Liu, Q., Davis, M. W., Hollopeter, G., Thomas, N., Jorgensen, N. B.,
microscopy of intact cells, tissues, and organisms. PLoS Biol. 9:e1001041. doi: et al. (2013). Ultrafast endocytosis at Caenorhabditis elegans neuromuscular
10.1371/journal.pbio.1001041 junctions. Elife 2, e00723. doi: 10.7554/eLife.00723
Shui, B., Ozer, A., Zipfel, W., Sahu, N., Singh, A., Lis, J. T., et al. (2012). RNA Watanabe, S., Punge, A., Hollopeter, G., Willig, K. I., Hobson, R. J., Davis, M. W.,
aptamers that functionally interact with green fluorescent protein and its et al. (2011). Protein localization in electron micrographs using fluorescence
derivatives. Nucleic Acids Res. 40, e39. doi: 10.1093/nar/gkr1264 nanoscopy. Nat. Methods 8, 80–84. doi: 10.1038/nmeth.1537
Siegel, M. S., and Isacoff, E. Y. (1997). A genetically encoded optical probe of Watanabe, S., Trimbuch, T., Camacho-Perez, M., Rost, B. R., Brokowski, B.,
membrane voltage. Neuron 19, 735–741. doi: 10.1016/S0896-6273(00)80955-1 Sohl-Kielczynski, B., et al. (2014). Clathrin regenerates synaptic vesicles from
Sochacki, K. A., Shtengel, G., van Engelenburg, S. B., Hess, H. F., and endosomes. Nature 515, 228–233. doi: 10.1038/nature13846
Taraska, J. W. (2014). Correlative super-resolution fluorescence and metal- Watkins, R. E. J., Rockett, P., Thoms, S., Clampitt, R., and Syms, R. (1986). Focused
replica transmission electron microscopy. Nat. Methods 11, 305–308. doi: ion-beam milling. Vacuum 36, 961–967. doi: 10.1016/0042-207x(86)90148-X
10.1038/nmeth.2816 White, J. G., Southgate, E., Thomson, J. N., and Brenner, S. (1986). The structure
Sonomura, T., Furuta, T., Nakatani, I., Yamamoto, Y., Unzai, T., Matsuda, W., et al. of the nervous system of the nematode Caenorhabditis elegans. Philos. Trans. R.
(2013). Correlative analysis of immunoreactivity in confocal laser-scanning Soc. Lond. B Biol. Sci. 314, 1–340. doi: 10.1098/rstb.1986.0056
microscopy and scanning electron microscopy with focused ion beam milling. Zhang, H., and Liu, K. K. (2008). Optical tweezers for single cells. J. R. Soc. Interface
Front. Neural Circuits 7:26. doi: 10.3389/fncir.2013.00026 5, 671–690. doi: 10.1098/rsif.2008.0052
Stauffer, T. P., Ahn, S., and Meyer, T. (1998). Receptor-induced transient reduction Zochowski, M., Wachowiak, M., Falk, C. X., Cohen, L. B., Lam, Y. W., Antic, S.,
in plasma membrane PtdIns(4,5)P2 concentration monitored in living cells. et al. (2000). Imaging membrane potential with voltage-sensitive dyes. Biol. Bull.
Curr. Biol. 8, 343–346. doi: 10.1016/S0960-9822(98)70135-6 198, 1–21. doi: 10.2307/1542798
Svitkina, T. (2009). Imaging cytoskeleton components by electron microscopy.
Methods Mol. Biol. 586, 187–206. doi: 10.1007/978-1-60761-376-3_10 Conflict of Interest Statement: The authors declare that the research was
Svitkina, T. M., Verkhovsky, A. B., and Borisy, G. G. (1995). Improved procedures conducted in the absence of any commercial or financial relationships that could
for electron microscopic visualization of the cytoskeleton of cultured cells. be construed as a potential conflict of interest.
J. Struct. Biol. 115, 290–303. doi: 10.1006/jsbi.1995.1054
Szczesny, P. J., Walther, P., and Muller, M. (1996). Light damage in rod outer Copyright © 2016 Begemann and Galic. This is an open-access article distributed
segments: the effects of fixation on ultrastructural alterations. Curr. Eye. Res. under the terms of the Creative Commons Attribution License (CC BY). The use,
15, 807–814. doi: 10.3109/02713689609017621 distribution or reproduction in other forums is permitted, provided the original
Tanasugarn, L., McNeil, P., Reynolds, G. T., and Taylor, D. L. (1984). author(s) or licensor are credited and that the original publication in this journal
Microspectrofluorometry by digital image processing: measurement of is cited, in accordance with accepted academic practice. No use, distribution or
cytoplasmic pH. J. Cell Biol. 98, 717–724. doi: 10.1083/jcb.98.2.717 reproduction is permitted which does not comply with these terms.
Advanced Fluorescence
Protein-Based Synapse-Detectors
Hojin Lee 1,2† , Won Chan Oh 1† , Jihye Seong 2,3 * and Jinhyun Kim 1,2 *
1
Center for Functional Connectomics, Korea Institute of Science and Technology, Seoul, South Korea, 2 Neuroscience
Program, Korea University of Science and Technology, Daejeon, South Korea, 3 Center for Diagnosis Treatment Care of
Dementia, Korea Institute of Science and Technology, Seoul, South Korea
by Santiago Ramon y Cajal (Nobel Laureate 1906). Since then, kinase II (CaMKII; Shen et al., 2000), α-amino-hydroxy-5-
the synapse as a structural and functional communication unit methyl-4-isoxazolepropionic acid receptor (AMPAR; Zamanillo
has been in the spotlight of neuroscientific inquiry (Cowan et al., et al., 1999) and so forth, that had been tagged with FPs.
2001). In fact, many studies have demonstrated that synaptic Recently, sophisticated molecular engineering has allowed even
events, such as changes in molecular composition, structure, more precise and detailed visualization of synaptic structure,
efficacy, and potentiation, play important roles in brain functions composition, and physiology (Chen et al., 2014; Fortin et al.,
including memory formation, perception, and other complex 2014; Figure 1, Table 1).
behaviors (Tsien et al., 1996; Markram et al., 1997; Malinow
and Malenka, 2002; Russo et al., 2010; Caroni et al., 2014). An
important focus has been to visualize the synapse and to measure pH-Sensitive FPs for Visualizing Vesicle
its activity. In 1954, DeRobertis and Palay first observed synapses Release: pHluorin, pHTomato, and pHuji
independently by electron microscopy (EM) and George Gray Although previous, straightforward, FP-based detection of
suggested there may be different types of synapses, i.e., excitatory synaptic distribution revealed many important details of synaptic
and inhibitory (De Robertis and Bennett, 1955; Palay and Palade, physiology, synaptic vesicle release/recycling and neural activity-
1955; Palay, 1956; Gray, 1959). EM provides enough resolution driven changes in membrane-bound synaptic proteins are
for nanometer-scale imaging of the synaptic ultrastructure, difficult to be detected by regular FP-tagging. Special sensors
something that cannot be achieved by light microscopy (LM) of vesicle secretion and neurotransmission have been developed
because of its diffraction limit. However, despite recent advances by linking vesicle membrane proteins with pH-sensitive mutants
that have reduced the time needed for image acquisition and of GFP called pHlourin (Miesenböck et al., 1998). The
reconstruction, EM remains inherently time-consuming, labor- fluorescence intensity of pHluorin largely depends on the
intensive, and volume-limited for large neural circuits. pH of its biochemical environment: in acidic environments
With the recent engineering of fluorescent proteins (FPs) with pH below 6.5, pHluorin is mostly nonfluorescent in
and new developments in light-favoring tissue clearing and 480 nm of light illumination, but becomes highly fluorescent
advanced optical methods, LM is rising as a potent alternative in neutral environments with pH around 7.4. This special
tool for investigating individual synapses in the context of feature of pHluorin was achieved by several mutations on
neural networks (Gray, 1959; Keller et al., 2008; Kim et al., residues important for the proton-relay of tyrosine 66 in the
2012; Tomer et al., 2012; Chung et al., 2013; Richardson chromophore (S147D, N149Q, T161I, S202F, and Q204T). These
and Lichtman, 2015). Imaging the synapse with LM by using amino acid substitutions, which set the pKa of pHluorin to
newly engineered FPs tagged to synaptic proteins or targeted around 7.0, can facilitate the pH-dependent switching of the
to synaptic structures enables fine-resolution illumination of electrostatic environment of the chromophore, allowing the pH-
synaptic anatomy and function in large neural circuits, possibly dependent changes in fluorescent intensity (Sankaranarayanan
in real-time. In this review, we describe recent FP-based and Ryan, 2000). When pHluorin is fused to presynaptic
molecular tools for imaging individual synapses and synaptic vesicle proteins such as VAMP2 (Miesenböck et al., 1998),
connectivity in the contexts of single- and dual component synaptophysin (Zhu et al., 2009), and vesicular glutamate
synaptic detection. We also identify crucial technologies: gene transporter (vGluT; Voglmaier et al., 2006), the release and
delivery of molecular synapse detector; tissue clearing for recycle of synaptic vesicles can be monitored as the pH inside
whole-brain imaging; and computational analysis, whose parallel the synaptic vesicles (∼5.5) transitions to the pH of the
development has potential to bridge synaptic sensor engineering extracellular environment (∼7.4). Thus, by tracking changes
and systems neuroscience. We end by proposing a scheme in pHluorin fluorescence intensity, one can detect real-time
of technological integration for synaptic neuroscience at the presynaptic exocytosis in living, active neurons. Similarly,
systems level. postsynaptic endo-/exo-cytosis and related receptor dynamics
can be visualized by pHluorin-fused mGluRs, for instance
SINGLE COMPONENT SYNAPTIC (Pelkey et al., 2007).
DETECTION More recently, red pH-sensitive FPs such as pHTomato (Li
and Tsien, 2012) and pHuji (Shen et al., 2014) have been
The discovery of green fluorescent protein (GFP) and its developed, allowing for the simultaneous monitoring of multiple
derivatives revolutionized the visualization of biological synaptic activities when combined with the green GFP-based
phenomena, including the individual synapse and its functions. sensors (e.g., GCaMP). pHTomato (pKa ∼ 7.8) was derived from
GFP and other FPs are relatively inert and small (27 kDa) and mStrawberry by introducing six mutations (F41T, F83L, S182K,
can be used to tag synaptic proteins while minimally interfering I194K, V195T, and G196D). Synaptophysin-fused pHTomato
with their normal functions. In fact, the distribution, trafficking, (sypHTomato) has been shown to be suitable for simultaneously
and physiological changes of synaptic proteins caused by neural monitoring the fusion of synaptic vesicles and, when paired
activity became evident in the last two decades, largely through with GCaMP3, postsynaptic Ca2+ changes in living neurons.
observations of synaptic proteins, such as synaptophysin (Li In addition, multiple synaptic events have been successfully
and Murthy, 2001) vesicle-associated membrane protein 2 measured by using sophisticated combinations of these pH-
(VAMP2; Ahmari et al., 2000), postsynaptic density protein-95 sensitive FP-fused synaptic proteins and spectrally distinct
(PSD-95; Nelson et al., 2013) calcium/calmodulin-dependent variants of optogentic stimulators (ChR2-T2A-vGluT-pHlourin
FIGURE 1 | Scheme of single component synapse detectors. (A) Graphical depiction of the conventional green fluorescent protein (GFP) tagging scheme and
summary of the expressed carrier-sensor construct. GFP (green circle) and other fluorescent proteins (FPs) are directly tagged to synaptic proteins in the presynaptic
terminal or postsynaptic spines. Major targets for tagging include presynaptic vesicle proteins (e.g., vesicle-associated membrane protein 2 (VAMP2) and
synaptophysin), postsynaptic receptors (e.g., α-amino-hydroxy-5-methyl-4-isoxazolepropionic acid receptor, AMPAR) and postsynaptic density protein-95 (PSD-95).
(B) pH-sensitive FP mutants are fused to synaptic vesicle membrane proteins, such as VAMP2, to visualize vesicle secretion/recycling and neurotransmission. A
pH-sensitive GFP variant, pHluorin does not fluoresce (gray circle) when inside the acidic chemical environment of the synaptic vesicle, but becomes highly
fluorescent (green circle) when the vesicle is released and exposed to the neutral extracellular environment. (C) Inhibition of Synaptic Release with CALI (InSynC):
attached to target SNARE proteins, molecular actuators such as mini small singlet oxygen generator (miniSOG; light blue filled-in circle) selectively inactivate specific
synaptic proteins that regulate vesicle release and other synaptic events. When illuminated with blue light, miniSOG stimulates generation of reactive oxygen species
(small, dark blue filled-in circles), which then oxidizes susceptible amino acid residues in target vesicle proteins and deactivates protein functions. (D) TimeSTAMP
effectively tracks spatiotemporally controlled protein synthesis and trafficking in living neurons. In the presence of a membrane-permeable protease inhibitor, NS3
protease (gray oval) activity is inhibited, and Venus C-terminus (Venus CT) and Venus N-terminus (Venus NT) reconstitute as fluorescent Venus (yellow circle).
Reconstituted Venus accumulates in the postsynaptic spine, the trafficking destination of the fused PSD-95. When the protease inhibitor is present, however, NS3
protease cleaves the protease target sites (gray circles flanking NS3), preventing Venus reconstitution.
and VChR1-T2A-synpHTomato). However, the pH-sensitivity Inhibition of Synaptic Release with CALI
of pHTomato is relatively low (3-fold change in pH 5.5–7.5). (InSynC)
Thus, a new red pH-sensitive FP, named pHuji, has been more Beyond merely visualizing the distributions and endo-/exocytosis
recently derived from mApple (Shaner et al., 2008) by including of synaptic proteins by tagging them with FPs and their pH-
a K163Y mutation, resulting in high pH sensitivity (20-fold sensitive variants, genetically encoded chromophore-assisted
change in pH 5.5–7.5). The use of different colored pH-sensitive light inactivation (CALI) has been developed to selectively
FPs together with a Ca2+ indicator and/or a spectrally distinct inactivate specific synaptic protein functions that regulate
optogenetic modulator offers a new promising readout system synaptic events such as synaptic release (Lin et al., 2013). CALI
for complex, coordinated synaptic events. is based on light-induced generation of reactive oxygen and the
consequent inactivation of nearby attached synaptic proteins. chosen because it is small (19 kDa) and monomeric, specific
Its original agents were synthetic chromophores for example to its substrate, and not cytotoxic, and most of all, NS3
malachite green (Jay, 1988), fluorescein (Beck et al., 2002), FlAsH shows high selectivity and efficiency of its cell-permeable
(Marek and Davis, 2002), ReAsH (Tour et al., 2003), and eosin inhibitor.
(Takemoto et al., 2011) and FPs such as KillerRed (Bulina et al., Although the first version of TimeSTAMP has worked
2006). To precisely inhibit synaptic release with improved target successfully in primary hippocampal neurons to track PSD-
specificity and inactivation efficiency compared to these CALI 95 accumulation during synaptic growth, and in Drosophila
agents, inhibition of synaptic release with CALI (InSynC) has for whole-brain mapping of newly synthesized CaMKII,
been recently developed using a newly engineered flavoprotein it required post hoc immunostaining against epitope tags,
called mini small singlet oxygen generator (miniSOG), fused limiting the benefits of pulse-chase labeling through time.
with the SNARE proteins VAMP2 and synaptophysin (Lin Thus, TimeSTAMP2 has been introduced, replacing HA-
et al., 2013). miniSOG was originally derived from the light, tag with split-fluorescence proteins (e.g., Yellow fluorescent
oxygen, and voltage 2 (LOV2) domain of phototropin, a protein, YFP; Orange fluorescent protein, OFP). In fluorescent
blue light photoreceptor (Shu et al., 2011). Under blue light TimeSTAMP2, for instance, Venus yellow fluorescent protein
illumination, this photoreceptor binds to and excites flavin is separated by an NS3 domain flanked by NS4A/B target
mononucleotide (FMN), which then functions as an oxygen sites into VenusNT (1-158 aa) and VenusCT (159-238 aa)
generator in cells. miniSOG contains the single amino acid for drug-dependent fluorescence. In the absence of BILN-
substitution of FMN-binding residue Cys426 to Gly on the LOV2 2061, VenusCT is cleaved and degraded by the active NS
domain, allowing for more efficient energy transfer to FMN, domain resulting in no yellow fluorescence, while VenusNT
and contains further mutations for enhanced brightness. When and CT are reconstituted as fluorescent forms after drug
illuminated by blue light, synaptic proteins can be selectively application. This allows optical pulse labeling of synaptic proteins
inactivated by miniSOG-mediated oxidization of susceptible such as PSD-95 and Neuroligin with a drug-defined temporal
residues such as tryptophan, tyrosine, histidine, cysteine, and resolution. Additionally, photo-oxidizing TimeSTAMP using
methionine. InSynC by miniSOG can selectively inhibit vesicle miniSOG inserted into TS:YFP can be used to visualize new
release at individual synapses in vitro and in vivo thanks proteins at an EM-based ultrastructural level (Butko et al.,
to the high efficiency of its light-induced oxygen generation, 2012).
independence of exogenous cofactors, and small size (106
residues, 14 kDa; Lin et al., 2013). However, further engineering DUAL COMPONENT SYNAPTIC
is required for investigating the functional dynamics of synaptic
DETECTION
circuits with physiologically relevant temporal resolutions, as
inactivation by InSynC persists relatively long (∼1 h) after light Thus far, we have described methods for detecting
stimulation. synapses by labeling single components, either pre- or
post-synaptic. Although these single-component tools
allow for detecting, measuring, and manipulating synaptic
Time-Specific Tagging for the Age structures and activities, they do not address the critical
Measurement of Proteins (TimeSTAMP) fact that synapses are bilateral microstructures involving
Another new strategy beyond merely visualizing synaptic both presynaptic terminal and postsynaptic density. For
proteins by tagging with FPs is time-specific tagging for reliable synapse detection, therefore, several recent studies
the age measurement of proteins (TimeSTAMP; Lin et al., introduced new methods for labeling synaptic interactions
2008). As spatiotemporally controlled protein synthesis between pre- and post-synaptic components, such as
and trafficking are critical for synaptogenesis, synaptic using split-FPs or enzymes to label neurexin-neuroligin
connectivity, and long-lasting changes in synapses, TimeSTAMP interactions. Here we review dual-component methods using
is beneficial for tracking important, newly synthesized synaptic handshaking-like transmembrane molecular interaction
proteins such as PSD-95 and CaMKII in living neurons. across the synaptic cleft, with particular attention to GFP
TimeSTAMP is a drug-controllable, time-specific tagging Reconstitution Across Synaptic Partners (GRASP; Figure 2,
strategy based on the hepatitis C virus NS3 protease and Table 1).
its cell-permeable inhibitor BILN-2061. PSD-95-GFP, for
example, was fused to NS3 protease flanked by NS4A/B
target sites and the C-terminal HA tag (PSD-95-GFP-TS- Mammalian GFP Reconstitution Across
HA). In the absence of the specific inhibitor BILN-2061, Synaptic Partners (mGRASP)
NS3 protease cleaves the NS4A/B target sites allowing the To detect particular synaptic connections with LM, the
C-terminal HA-tag to be cleaved from PSD-95-GFP and Mammalian GFP Reconstitution Across Synaptic Partners
degraded; but drug application will allow the HA-tag to (mGRASP) technique takes advantage of the complementarity
be accumulated. This allows distinguishing between newly of two non-fluorescent split-GFP fragments, each of which is
and previously synthesized PSD-95 by measuring HA/GFP tethered specifically to the pre- and postsynaptic membrane,
signals at a time defined by the drug application. For respectively (Kim et al., 2012). When two neurons, each
the TimeSTAMP strategy, the NS3 protease domain was expressing one of the fragments, are closely opposed across a
FIGURE 2 | Scheme of dual component synapse detectors. (A) GFP reconstitution-dependent molecular tools detect pre- and postsynaptic membrane
apposition through the reconstitution of spGFP fragments each attached to the extracellular ends of membrane protein carriers. Synapse-targeted GFP
Reconstitution Across Synaptic Partners (GRASP) utilizes a presynaptic component of spGFP11 (spG11, circular wedge) linked to a CD4-neurexin-1β C-terminus
(NxCT) fusion through flexible peptide linkers (gray zigzag), and a postsynaptic component of spGFP1-10 (spG1-10) and truncated neuroligin-1 (tNg) connected with
a linker. SpG11 and spG1-10 unite at synapses, not random membrane contacts. In activity-dependent X-RASP the presynaptic split-XFP1-10 (spX1-10) is fused to
SNARE proteins such as synaptobrevin (syb), and unites with CD4-bound spG11 when presynaptic activity triggers vesicle release. (B) Neurexin (Nx)-neuroligin (Ng)
interaction-dependent molecular tools detect individual synapses through fluorescent protein (FP)-based (SynView) and enzymatic reaction-based
(Interaction-Dependent Probe Incorporation Mediated by Enzymes, ID-PRIME) direct visualization of Nx-Ng interaction. In SynView, spG11 is inserted between the
esterase and proximal extracellular domain of Ng, and spG1-10 in the middle of the LNS domain of Nx. GFP reconstitutes when Nx and Ng unites. ID-PRIME uses
the identical set of synaptic protein carriers, but tandem LAP tags (3xLAP) and lipoic acid ligase A (LplA) are attached to C-terminus of Ng esterase domain and Nx
LNS domain, respectively. Nx-Ng interaction initiates LplA-mediated lipoic acid tagging to 3xLAP, and signals useful for imaging originate from antibody-fluorophore
conjugates.
synaptic cleft, the split fragments are reconstituted as fluorescent approximately 20 nm-wide synaptic cleft without gross changes
GFP in that location. This dual component synapse detection in endogenous synaptic organization and physiology. Based
bypasses the Abbe’s diffraction limit and allows relatively on published sequences from NCBI, we generated synapse-
rapid and accurate synapse mapping with LM. This method specific chimeric carriers. Both the pre- and postsynaptic
takes advantage of the fast folding kinetics and stability after mGRASP components share a common six-subcomponent
maturation of superfolder GFP (Pédelacq et al., 2006), split framework: (1) a signal peptide; (2) a split-GFP fragment;
into two fragments, namely spGFP1-10 (first 214 residues (3) an extracellular domain; (4) a transmembrane domain;
comprising ten β barrels) and spGFP11 (16 residues, 11th β- (5) an intracellular domain; and (6) a fluorescent protein for
barrel strand). The GRASP technique was initially implemented neurite and soma visualization. The presynaptic mGRASP
in C. elegans (Feinberg et al., 2008; Park et al., 2011) and component consists of the signal peptide of nematode β-
later in Drosophila (Gordon and Scott, 2009; Fan et al., 2013; integrin (PAT-3, residues 1–29), GFP β-strand 11 (spGFP11,
Gorostiza et al., 2014). In the original GRASP constructs, 16 residues), two flexible GGGGS linkers, the extracellular
membrane carriers (human CD4) for split-GFP fragments were and transmembrane domains of human CD4-2 (residues
not synapse-specific—fluorescence could arise wherever any 25–242). Rat neurexin-1β (residues 414–468) containing
membranes expressing fragment pairs were closely apposed. the PDZ-binding motif constitutes the intracellular domain
This became a critical issue for mammalian synapse detection for maintenance of correct localization at the presynaptic
because the mammalian nervous system contains much more site. mCerulean is fused to the intracellular end of the
compactly intermingled neurites than the invertebrate nervous presynaptic mGRASP to visualize axonal expression. The
system. postsynaptic mGRASP component is based primarily on
In a previous study, we engineered and successfully applied mouse neuroligin-1, which interacts with presynaptic adhesion
mGRASP for mapping fine-scale synaptic connectivity in proteins including β-neurexins, and mediates the formation
the mouse brain (Kim et al., 2012). The primary features of and maintenance of synapses between neurons. To prevent
mGRASP include targeting specific synapses, and matching the nonspecific synaptogenesis through interactions with its
Detection/working pH-sensitive FPs: pHluorin, miniSOG- Reconstitution of Reconstitution of Reconstitution of Reconstitution of Reconstitution of Enzymatic
module pHTomato generated reactive VenusNT and spGFP1-10 and spGFP11 spGFP1-10 and spXFP1-10 spGFP1-10 interaction of LpIA
oxygen VenusCT spGFP11 and spXFP11 and spGFP11 and LAP
Synaptic targeting Synaptophysin, VAMP2, Synaptophysin, PSD-95, CaMKII, Human CD4, PTP-3A, Rat neurexin1-β, Synaptobrevin, Rat neurexin-1β, Human
Model Systems
In vitro HEK, primary cultured Primary cultured HEK, primary N/A N/A HEK HEK, COS-7, primary HEK, primary
neurons, organotypic slices neurons, cultured neurons cultured neurons cultured neurons
organotypic slices
In vivo N/A N/A N/A C.elegans, Drosophila Mouse Drosophila N/A N/A
Remarks Tracking of presynaptic Optically inhibiting Optical pulse-chase in vivo synaptic detection in in vivo synaptic Activity-dependent Synapse labeling by Enzyme-based
vesicle release, synaptic release labeling of synaptic invertebrate circuits detection in vertebrate labeling of multiple direct interaction of visualization of
postsynaptic receptor with good spatial proteins with high circuits synaptic inputs neurexin-neuroligin direct interaction of
trafficking resolution spatiotemporal neurexin-neuroligin
resolution
Lacks in vivo application Lacks Awaits mammalian Limited to apply Limited to detect Limited to apply Lacks in vivo Requires separate
with quantitative analysis physiologically in vivo application mammalian system functional synaptic mammalian system, verification and introduction of
relevant temporal connections limited temporal application exogenous ligase
resolutions due to resolutions and antibody-
slow recovery fluorophore
conjugates
References Miesenböck et al. (1998), Lin et al. (2013) Lin et al. (2008) Feinberg et al. (2008), Feng et al. (2012), Karuppudurai et al. Tsetsenis et al. Liu et al. (2013)
Sankaranarayanan and and Butko et al. Gordon and Scott (2009), Kim et al. (2012) (2014), (2014)
Ryan (2000), (2012) Gong et al. (2010), and Druckmann et al. Frank et al. (2015),
Kopec et al. (2006), Park et al. (2011), (2014) Macpherson et al. (2015)
Voglmaier et al. (2006) Yuan et al. (2011), and Li et al. (2016)
and Li and Tsien (2012) Fan et al. (2013),
Pech et al. (2013)
and Gorostiza et al. (2014)
endogenous partner, neurexin, the extracellular esterase 2-photon imaging together with Ca2+ indicators and for
domain of neuroligin is deleted in the main skeleton of the fine-scale synapse labeling from multiple inputs within the
postsynaptic mGRASP. Similar to presynaptic mGRASP, single neuron, because it has proved difficult to separate
post-mGRASP is composed a signal peptide from the CFP/GFP/YFP signals in the complex mammalian nerve
esterase-truncated neuroligin-1 (residues 1–49), GFP β- system.
strand 1-10 (spGFP1-10, 648 residues), the extracellular,
transmembrane, and intracellular regions of neuroligin (71, Engineering Carriers
19, and 127 residues, respectively), followed by the self- for Activity-Dependent GRASP
cleavable 2A peptide-fused dTomato for visualizing postsynaptic To understand functional organizations underlying complex
neuronal morphology. This optimized mGRASP enabled the brain functions that go beyond structural connectivity patterns,
comprehensive synaptic connectivity mapping of hippocampal it will be essential to identify active synaptic connectivity at
CA3-CA1 and identified spatially structured patterns of synaptic defined times and conditions. For the activity-dependent GRASP
connectivity (Druckmann et al., 2014). It is important to note system, synaptobrevin (syb), a key constituent of synaptic
that brain-wide synapse detection for comprehensive fine-scale vesicle membrane, was used as a synaptic carrier instead of
mapping becomes achievable not only by advanced molecular constantly membrane-targeted carriers in the previous GRASP
engineering to label the synapse such as mGRASP, but also by systems. The straightforward fusion of syb and spGFP1-10
appropriately engineered computational analysis (Feng et al., (syb:spGFP1-10) with the original CD4:spGFP11 can together
2012, 2015). form activity-dependent GRASP, called the syb:GRASP system.
Current GRASP technology has proved to be a tool suitable Given activity-dependent interactions of syb with SNARE-SM
for rapid and accurate mapping of synaptic connectivity in protein complex triggering synaptic vesicle release, syb:GRASP
nematode (Feinberg et al., 2008), fruit fly (Gordon and Scott, showed preferential labeling of active synapses as a boost of
2009), and mouse (Kim et al., 2012; Druckmann et al., 2014). Yet, GRASP fluorescence signals in well-studied thermosensory and
further improvements to GRASP such as multi-colored FPs for olfactory circuits in the fruit fly (Macpherson et al., 2015). This
analyzing convergent synaptic inputs, various carriers for neural new strategy for mapping active synaptic connectivity might
activities, and tailored computational analyses will provide a clear expedite the mapping of functional connectivity at the synapse
overview of complex synaptic connectivity and its operation. We level in a way previously achieved only by difficult combinations
next discuss recent efforts in these directions. of Ca2+ imaging and exhaustive EM reconstruction (Bock
et al., 2011; Briggman et al., 2011). Yet, these new techniques
Engineering FPs for Multi-Color GRASP raise basic concerns about possible side effects caused by
A neuron oftentimes receives multiple inputs from different the overexpression of key synaptic vesicle proteins, and need
presynaptic neurons of distinct cell types and/or various further optimization before they can be applied to mammalian
brain areas. For comprehensive mapping of complex synaptic networks.
circuits, multiple synaptic detection is beneficial, as described
earlier in the section describing pH-sensitive FPs. The
current version of GRASP restricts convergent synaptic FP- and Enzyme-Based Visualization
mapping because it relies on only a single pair of spGFP of Neuroligin-Neurexin Interaction
fragments such that spectral overlap of its signals hinders An alternative way to detect synapses has been suggested
simultaneous imaging with previously well-established GFP- by imaging neurexin-neuroligin interactions. Presynaptic
based tools such as GCaMP. Therefore, the multi-color neurexin and postsynaptic neurolign are transmembrane
GRASP (XRASP) components have been developed recently adhesion proteins and their interaction at the synaptic
(Macpherson et al., 2015; Li et al., 2016). Given that Cyan cleft has been believed to be a key process for synaptic
fluorescent protein (CFP), GFP, and YFP have identical formation, maintenance, and connectivity (Li and Sheng,
structures except for several residues in the chromophore 2003; Graf et al., 2004; Chen et al., 2010; Krueger et al.,
located in beta barrels of 1-10, C-RASP and Y-RASP were 2012). Therefore, it is thought that identifying sites of
generated by color-shifting mutations in spGFP1-10 (Y66W, neurexin-neuroligin interactions could provide selective
S72A, F145A∗ , N146I, and H148D for C-RASP, ∗ additional visualization of synapses. Similar to the GRASP approaches,
mutation in Li et al., 2016, T65G, V68L, S72A and T203Y this method, based on neuroligin-neurexin interactions using
for Y-RASP) paired with the unaltered spGFP11 (Li et al., splitGFP, is called SynView. It is composed of spGFP1-
2016). Multi-color GRASP (XRASP) was tested in vivo in 10-inserted neurexin-1β between residues N275-D276 or
several circuits, including the Kenyon cells of the mushroom A200-G201 for the presynaptic component, and the split-
body, and projection neurons of the thermosensory and GFP11-inserted neuroligin-1 in the C-terminal end of the
olfactory systems in the fruit fly. Reconstructed CFP and YFP extracellular esterase (between Q641-Y642) for the postsynaptic
signals showed minimal spectral overlap with the GRASP component (Tsetsenis et al., 2014). Another method for
emission and excitation spectra. Extended choices of multi- visualizing neuroligin-neurexin interactions is based on
color GRASP (XRASP) will allow simultaneous imaging of mutated bacterial lipoic acid ligase (LpIA) and lipoic acid
multiple, convergent connectivity and functional activity. acceptor peptide (LAP) tags (Uttamapinant et al., 2010, 2012),
However, more red-shifted XRASP is required for in vivo and is called Interaction-Dependent Probe Incorporation
Mediated by Enzymes (ID-PRIME; Liu et al., 2013). Using of Cre recombinase. When this PSD-95-ENABLED mouse
the same principle of SynView, ID-PRIME was designed to line is crossed with a dopaminergic cell-type specific DAT-
detect trans-synaptic contacts of neurexin and neuroligin Cre line, for instance, PSD-95mVenus is expressed specifically
enzymatically using lipoic acid. These two methods successfully in dopaminergic neurons. PSD-95-ENABLED was shown to
imaged the direct interactions of neurexin-neuroligin synaptic functionally replace wild-type PSD-95 and to sparsely label
adhesive molecules. However, these methods seem restricted a particular cell-type, allowing insights into the detailed
to particular investigations of the dynamics of neurexin- distribution and dynamics of synapses. In applying this
neuroligin, rather general synaptic mapping. In addition, strategy to a wide variety of synaptic proteins, one concern
it is known that overexpressing these synaptic adhesion with using the ENABLED strategy is that it is limited to
molecules causes substantial structural and physiological synaptic proteins that are compatible with C-terminal FP-
perturbations of normal synaptic compartments and cleft tagging.
structures. Thanks to incredibly fast developments and improvements in
genetic manipulation technologies such as clustered regularly-
BRIDGING TOOLS BETWEEN interspaced short palindromic repeats (CRISPR) and effector
MOLECULAR SYNAPSE DETECTORS nucleases Cas system (Cong et al., 2013; Wang et al., 2013;
AND BRAIN-WIDE SYNAPSE MAPPING Fujii et al., 2014; Aida et al., 2015), the generation of synapse
detector KI mouse lines is becoming time- and cost-efficient, and
Thus far, we have described methods for detecting synapses is dramatically facilitating synapse mapping.
focusing on molecular engineering. To exert these genetically
encoded synapse detectors to brain-wide synaptic mapping Viral System-Based Gene Delivery
at the system level, there need to be critically partnered Spatially targeted gene delivery into the mature brain can
technologies such as gene delivery, advanced imaging, and digital be achieved through stereotactic microinjection of viral
representation platform that are appropriate for neural system vectors expressing FP-tagged proteins. Diverse virus families
(Figure 3). have been recruited into this effort: Retroviridae (e.g.,
lentivirus), Parvoviridae (e.g., rAAV), Adenoviridae (e.g.,
Gene Delivery System for Synapse canine adenovirus), and Alphaviridae (e.g., sindbis virus),
including some that cross the synapse, such as Rhabdoviridae
Detectors in the Mouse Brain (e.g., rabies virus) and Herpesviridae (e.g., HSV-1 and
Improving the targeting specificity for types of cells and
pseudorabies virus; Nassi et al., 2015). Because of their low
scaling up the scope of synaptic sensor delivery are among
cytotoxicity and stable expression, lentiviruses and rAAVs
the crucial initial steps needed to expand single-synapse
are the most widely used viral vectors in neuroanatomical
level analysis to systems level neuroscience. Gene delivery
tracing studies and human clinical trials testing gene therapy.
technologies, particularly for the nervous system, have
In fact, spatially restricted injection of Cre-(in)dependent
developed at a breathtaking pace over the past decades,
rAAV vectors have been used for mGRASP-assisted synaptic
establishing methodologies that can be classified into
mapping.
two main categories: germ line manipulation and viral
In parallel, ongoing efforts include searching for and
injection.
engineering new types of virus to allow transduction efficiency/
specificity and retrograde infection. Canine adenovirus has
Genetic Manipulation-Based Gene Delivery drawn attention because of its strong retrograde transport
To avoid side effects that can sometimes be caused by the capability and relatively large payload size (∼30 kb); further,
overexpression of FP-tagged synaptic proteins, and to mimic several successful applications of Canine adenovirus in the mouse
expression patterns of endogenous proteins, gene knock- brain suggest a powerful complement to the lentivirus and rAAV
in (KI) technology has been used to substitute wild-type (Bru et al., 2010; Ekstrand et al., 2014; Junyent and Kremer,
genes with FP-tagged copies. For synaptic detectors based 2015; Schwarz et al., 2015). Additionally, systemic delivery of
on universal synaptic proteins such as PSD-95, however, viral vectors in animal models has proved to be effective and
this standard KI method leads to expression of FP-tagged safe for various serotypes of AAV (Foust et al., 2009; Bevan
synapse detectors globally, throughout the brain, which makes et al., 2011; Gray et al., 2011; Yang et al., 2014; Deverman et al.,
it difficult to acquire high resolution images of a particular 2016).
cell type. Therefore, a recent study introduced a conditional
KI strategy called endogenous labeling via exon duplication
(ENABLED; Fortin et al., 2014). In the ENABLED strategy, Advanced Brain-Wide Imaging and Digital
a knocked-in gene cassette is composed of the floxed last Representation of Synaptic Detectors
exon and 3’UTR of PSD-95 followed by its mVenus-tagged Once FP-based synaptic detectors are introduced into the
duplicated last exon and 3’UTR. The FP-tagged duplicated brain as described above, appropriate brain-wide imaging and
gene will be selectively expressed only in Cre-expressing digital representation technologies are necessary to map synaptic
neurons because its translation is designed to be blocked by connectivity. Happily, there has been remarkable progress
translation stop signals in the endogenous copy in the absence in tissue clearing methodology such as BABB (Dodt et al.,
FIGURE 3 | Convergence of technologies for synaptic neuroscience at the systems level. (A) Synapse-detector genes can be delivered to target neural
components through germline manipulation-based and viral system-based methods. Exogenous genetic materials can be effectively introduced to the germline with
high expression specificity via Cre-mediated conditional knock-in (KI) strategies and the homology directed repair (HDR) pathway of the clustered
regularly-interspaced short palindromic repeats (CRISPR)/Cas system. In a Cre-mediated conditional KI strategy called endogenous labeling via exon duplication
(ENABLED), a KI cassette including duplicates of exons 19 and 20 and the 3’UTR of PSD-95 is generated. The first duplicate is flanked by head-to-tail oriented loxP
sites (black arrows), while the second duplicate has monovalent Venus (mV) inserted between exon 20 and the 3’UTR. The two duplicate sequences, along with
exon 18, are flanked by a set of homology arms (h arm), which mediates KI cassette insertion through homologous recombination. Cre-lox recombination excises the
first duplicate containing translation stop signals and polyadenylation sequences, and activates mV expression only in Cre-expressing neurons. Transient expression
of sensor genes using viral vectors benefits from efforts to engineer expression through Cre-dependent expression cassettes e.g., Cre-dependent Mammalian
GRASP (mGRASP) constructs and newly engineered serotypes that have more selective tropism and transduction efficiency (e.g., AAV-PHP.B). Local tissue or
systemic injection of such viral systems can lead to flexible, versatile gene delivery in mature organisms. (B) Combination of light-favoring brain clearing, whole-brain
imaging, and computational techniques for three-dimensional synapse mapping enables single-synapse level analysis of synaptic profiles across the whole brain.
Further improvements in lipid extraction, refractive index matching, advanced light-sheet microscopy, and large-scale data processing and 3D reference space
generation will accelerate systems neuroscience at the synaptic scale.
2007), CUBIC (Susaki et al., 2014) 3DISCO (Ertürk et al., sectioning. These new clearing techniques, together with
2012), immunolabeling-enabled three-dimensional imaging of advanced optical methods such as fluorescent selective plane
solvent-cleared organs (iDISCO; Renier et al., 2014), SeeDB illumination microscopy (fSPIM; Huisken et al., 2004), will
(Ke et al., 2013) and CLARITY (Chung et al., 2013; Tomer allow high-throughput whole brain imaging. Also, once images
et al., 2014) that are needed to prepare the intact brain are acquired, digital representation programs are necessary
for imaging while avoiding distortions caused by physical to reliably extract synaptic wiring information from the
images, and to bring data from different sections and animals brain will provide firm grounds for further anatomical and
into register with one another (Johnson et al., 2010; Oh functional studies and such mapping is essential for analyzing
et al., 2014). Improvements in this software will be greatly large-scale information processing phenomena. We believe that
beneficial. rapid, scalable synaptic cartography with the triad of single-
In our view, new clearing methods, optics, and tailored synapse resolution, brain-wide scope, and cell-type-specific
computational analysis platforms together with advanced connectivity requires a synergistic combination of advanced
synaptic detectors such as mGRASP are very promising technologies, and that FP-based neurosensors are the key to this
developments for the complete mapping of mammalian synaptic grand integrative project.
connectivity.
AUTHOR CONTRIBUTIONS
CONCLUSION AND PERSPECTIVE
HL, WCO, JS, and JK wrote this manuscript.
Here we reviewed new FP-based synapse detection techniques,
which are useful for rapidly imaging individual synapses and ACKNOWLEDGMENTS
synaptic connectivity in the whole brain with LM. Recent
technical developments have allowed a focus of neuroscience to HL, WCO, JK, and JS are supported by the Korea Institute of
move from individual synapses to ensemble interactions among Science and Technology (KIST) Institutional Program (2E26190
neurons of various cell types through multiple synaptic pathways. and 2N41660, respectively) and the Samsung Science and
Comprehensively mapping individual synapses in the whole Technology Foundation.
REFERENCES Chung, K., Wallace, J., Kim, S.-Y., Kalyanasundaram, S., Andalman, A. S.,
Davidson, T. J., et al. (2013). Structural and molecular interrogation
Ahmari, S. E., Buchanan, J., and Smith, S. J. (2000). Assembly of presynaptic active of intact biological systems. Nature 497, 332–337. doi: 10.1038/nature
zones from cytoplasmic transport packets. Nat. Neurosci. 3, 445–451. doi: 10. 12107
1038/74814 Cong, L., Ran, F. A., Cox, D., Lin, S., Barretto, R., Habib, N., et al. (2013). Multiplex
Aida, T., Chiyo, K., Usami, T., Ishikubo, H., Imahashi, R., Wada, Y., et al. (2015). genome engineering using CRISPR/Cas systems. Science 339, 819–823. doi: 10.
Cloning-free CRISPR/Cas system facilitates functional cassette knock-in in 1126/science.1231143
mice. Genome Biol. 16:87. doi: 10.1186/s13059-015-0653-x Cowan, W. M., Südhof, T. C., and Stevens, C. F. (2001). Synapses. Baltimore: JHU
Beck, S., Sakurai, T., Eustace, B. K., Beste, G., Schier, R., Rudert, F., et al. Press.
(2002). Fluorophore-assisted light inactivation: a high-throughput tool for De Robertis, E. D., and Bennett, H. S. (1955). Some features of the submicroscopic
direct target validation of proteins. Proteomics 2, 247–255. doi: 10.1002/1615- morphology of synapses in frog and earthworm. J. Biophys. Biochem. Cytol. 1,
9861(200203)2:3<247::aid-prot247 >3.0.co;2-k 47–58. doi: 10.1083/jcb.1.1.47
Bevan, A. K., Duque, S., Foust, K. D., Morales, P. R., Braun, L., Schmelzer, L., et al. Deverman, B. E., Pravdo, P. L., Simpson, B. P., Kumar, S. R., Chan, K. Y.,
(2011). Systemic gene delivery in large species for targeting spinal cord, brain Banerjee, A., et al. (2016). Cre-dependent selection yields AAV variants for
and peripheral tissues for pediatric disorders. Mol. Ther. 19, 1971–1980. doi: 10. widespread gene transfer to the adult brain. Nat. Biotechnol. 34, 204–209.
1038/mt.2011.157 doi: 10.1038/nbt.3440
Bock, D. D., Lee, W.-C. A., Kerlin, A. M., Andermann, M. L., Hood, G., Dodt, H.-U., Leischner, U., Schierloh, A., Jährling, N., Mauch, C. P., Deininger, K.,
Wetzel, A. W., et al. (2011). Network anatomy and in vivo physiology of visual et al. (2007). Ultramicroscopy: three-dimensional visualization of neuronal
cortical neurons. Nature 471, 177–182. doi: 10.1038/nature09802 networks in the whole mouse brain. Nat. Methods 4, 331–336. doi: 10.
Briggman, K. L., Helmstaedter, M., and Denk, W. (2011). Wiring specificity in 1038/nmeth1036
the direction-selectivity circuit of the retina. Nature 471, 183–188. doi: 10. Druckmann, S., Feng, L., Lee, B., Yook, C., Zhao, T., Magee, J. C., et al. (2014).
1038/nature09818 Structured synaptic connectivity between hippocampal regions. Neuron 81,
Bru, T., Salinas, S., and Kremer, E. J. (2010). An update on canine adenovirus type 629–640. doi: 10.1016/j.neuron.2013.11.026
2 and its vectors. Viruses 2, 2134–2153. doi: 10.3390/v2092134 Ekstrand, M. I., Nectow, A. R., Knight, Z. A., Latcha, K. N., Pomeranz, L. E., and
Bulina, M. E., Chudakov, D. M., Britanova, O. V., Yanushevich, Y. G., Friedman, J. M. (2014). Molecular profiling of neurons based on connectivity.
Staroverov, D. B., Chepurnykh, T. V., et al. (2006). A genetically encoded Cell 157, 1230–1242. doi: 10.1016/j.cell.2014.03.059
photosensitizer. Nat. Biotechnol. 24, 95–99. doi: 10.1038/nbt1175 Ertürk, A., Becker, K., Jährling, N. J. A., Mauch, C. P., Hojer, C. D., Egen, J. G., et al.
Butko, M. T., Yang, J., Geng, Y., Kim, H. J., Jeon, N. L., Shu, X., et al. (2012). (2012). Three-dimensional imaging of solvent-cleared organs using 3DISCO.
Fluorescent and photo-oxidizing TimeSTAMP tags track protein fates in Nat. Protoc. 7, 1983–1995. doi: 10.1038/nprot.2012.119
light and electron microscopy. Nat. Neurosci. 15, 1742–1751. doi: 10.1038/ Fan, P., Manoli, D. S., Ahmed, O. M., Chen, Y., Agarwal, N., Kwong, S., et al.
nn.3246 (2013). Genetic and neural mechanisms that inhibit Drosophila from mating
Caroni, P., Chowdhury, A., and Lahr, M. (2014). Synapse rearrangements upon with other species. Cell 154, 89–102. doi: 10.1016/j.cell.2013.06.008
learning: from divergent-sparse connectivity to dedicated sub-circuits. Trends Feinberg, E. H., VanHoven, M. K., Bendesky, A., Wang, G., Fetter, R. D., Shen, K.,
Neurosci. 37, 604–614. doi: 10.1016/j.tins.2014.08.011 et al. (2008). GFP reconstitution across synaptic partners (GRASP) defines cell
Chen, Y., Akin, O., Nern, A., Tsui, C. Y. K., Pecot, M. Y., and Zipursky, S. L. contacts and synapses in living nervous systems. Neuron 57, 353–363. doi: 10.
(2014). Cell-type-specific labeling of synapses in vivo through synaptic 1016/j.neuron.2007.11.030
tagging with recombination. Neuron 81, 280–293. doi: 10.1016/j.neuron.2013. Feng, L., Zhao, T., and Kim, J. (2012). Improved synapse detection for mGRASP-
12.021 assisted brain connectivity mapping. Bioinformatics 28, i25–i31. doi: 10.
Chen, S. X., Tari, P. K., She, K., and Haas, K. (2010). Neurexin-neuroligin cell 1093/bioinformatics/bts221
adhesion complexes contribute to synaptotropic dendritogenesis via growth Feng, L., Zhao, T., and Kim, J. (2015). neuTube 1.0: a new design for efficient
stabilization mechanisms in and in vivo. Neuron 67, 967–983. doi: 10.1016/j. neuron reconstruction software based on the SWC format. eNeuro 2, 1–10.
neuron.2010.08.016 doi: 10.1523/ENEURO.0049-14.2014
Fortin, D. A., Tillo, S. E., Yang, G., Rah, J. C., Melander, J. B., Bai, S., et al. (2014). Li, Z., and Murthy, V. N. (2001). Visualizing postendocytic traffic of synaptic
Live imaging of endogenous PSD-95 using ENABLED: a conditional strategy to vesicles at hippocampal synapses. Neuron 31, 593–605. doi: 10.1016/s0896-
fluorescently label endogenous proteins. J. Neurosci. 34, 16698–16712. doi: 10. 6273(01)00398-1
1523/JNEUROSCI.3888-14.2014 Li, Z., and Sheng, M. (2003). Some assembly required: the development of
Foust, K. D., Nurre, E., Montgomery, C. L., Hernandez, A., Chan, C. M., neuronal synapses. Nat. Rev. Mol. Cell Biol. 4, 833–841. doi: 10.1038/nrm1242
and Kaspar, B. K. (2009). Intravascular AAV9 preferentially targets neonatal Li, Y., and Tsien, R. W. (2012). pHTomato, a red, genetically encoded indicator
neurons and adult astrocytes. Nat. Biotechnol. 27, 59–65. doi: 10.1038/ that enables multiplex interrogation of synaptic activity. Nat. Neurosci. 15,
nbt.1515 1047–1053. doi: 10.1038/nn.3126
Fujii, W., Onuma, A., Sugiura, K., and Naito, K. (2014). Efficient generation of Lin, M. Z., Glenn, J. S., and Tsien, R. Y. (2008). A drug-controllable tag for
genome-modified mice via offset-nicking by CRISPR/Cas system. Biochem. visualizing newly synthesized proteins in cells and whole animals. Proc. Natl.
Biophys. Res. Commun. 445, 791–794. doi: 10.1016/j.bbrc.2014.01.141 Acad. Sci. U S A 105, 7744–7749. doi: 10.1073/pnas.0803060105
Frank, D. D., Jouandet, G. C., Kearney, P. J., Macpherson, L. J., and Gallio, M. Lin, J. Y., Sann, S. B., Zhou, K., Nabavi, S., Proulx, C. D., Malinow, R., et al.
(2015). Temperature representation in the Drosophila brain. Nature 519, (2013). Optogenetic inhibition of synaptic release with chromophore-assisted
358–361. doi: 10.1038/nature14284 light inactivation (CALI). Neuron 79, 241–253. doi: 10.1016/j.neuron.2013.
Gong, Z., Liu, J., Guo, C., Zhou, Y., Teng, Y., and Liu, L. (2010). Two pairs of 05.022
neurons in the central brain control Drosophila innate light preference. Science Liu, D. S., Loh, K. H., Lam, S. S., White, K. A., and Ting, A. Y. (2013). Imaging
330, 499–502. doi: 10.1126/science.1195993 trans-cellular neurexin-neuroligin interactions by enzymatic probe ligation.
Gordon, M. D., and Scott, K. (2009). Motor control in a Drosophila taste circuit. PLoS One 8:e52823. doi: 10.1371/journal.pone.0052823
Neuron 61, 373–384. doi: 10.1016/j.neuron.2008.12.033 Macpherson, L. J., Zaharieva, E. E., Kearney, P. J., Alpert, M. H., Lin, T.-Y.,
Gorostiza, E. A., Depetris-Chauvin, A., Frenkel, L., Pírez, N., and Ceriani, M. F. Turan, Z., et al. (2015). Dynamic labelling of neural connections in multiple
(2014). Circadian pacemaker neurons change synaptic contacts across the day. colours by trans-synaptic fluorescence complementation. Nat. Commun.
Curr. Biol. 24, 2161–2167. doi: 10.1016/j.cub.2014.07.063 6:10024. doi: 10.1038/ncomms10024
Graf, E. R., Zhang, X., Jin, S.-X., Linhoff, M. W., and Craig, A. M. (2004). Malinow, R., and Malenka, R. C. (2002). AMPA receptor trafficking and synaptic
Neurexins induce differentiation of GABA and glutamate postsynaptic plasticity. Annu. Rev. Neurosci. 25, 103–126. doi: 10.1146/annurev.neuro.25.
specializations via neuroligins. Cell 119, 1013–1026. doi: 10.1016/j.cell. 112701.142758
2004.11.035 Marek, K. W., and Davis, G. W. (2002). Transgenically encoded protein
Gray, E. G. (1959). Axo-somatic and axo-dendritic synapses of the cerebral cortex: photoinactivation (FlAsH-FALI): acute inactivation of synaptotagmin I.
an electron microscope study. J. Anat. 93, 420–433. Neuron 36, 805–813. doi: 10.1016/S0896-6273(02)01068-1
Gray, S. J., Matagne, V., Bachaboina, L., Yadav, S., Ojeda, S. R., and Samulski, R. J. Markram, H., Lübke, J., Frotscher, M., and Sakmann, B. (1997). Regulation of
(2011). Preclinical differences of intravascular AAV9 delivery to neurons and synaptic efficacy by coincidence of postsynaptic APs and EPSPs. Science 275,
glia: a comparative study of adult mice and nonhuman primates. Mol. Ther. 19, 213–215. doi: 10.1126/science.275.5297.213
1058–1069. doi: 10.1038/mt.2011.72 Miesenböck, G., De Angelis, D. A., and Rothman, J. E. (1998). Visualizing
Huisken, J., Swoger, J., Del Bene, F., Wittbrodt, J., and Stelzer, E. H. K. (2004). secretion and synaptic transmission with pH-sensitive green fluorescent
Optical sectioning deep inside live embryos by selective plane illumination proteins. Nature 394, 192–195. doi: 10.1038/28190
microscopy. Science 305, 1007–1009. doi: 10.1126/science.1100035 Nassi, J. J., Cepko, C. L., Born, R. T., and Beier, K. T. (2015). Neuroanatomy goes
Jay, D. G. (1988). Selective destruction of protein function by chromophore- viral!. Front. Neuroanat. 9:80. doi: 10.3389/fnana.2015.00080
assisted laser inactivation. Proc. Natl. Acad. Sci. U S A 85, 5454–5458. doi: 10. Nelson, C. D., Kim, M. J., Hsin, H., Chen, Y., and Sheng, M. (2013).
1073/pnas.85.15.5454 Phosphorylation of threonine-19 of PSD-95 by GSK-3β is required for PSD-95
Johnson, G. A., Badea, A., Brandenburg, J., Cofer, G., Fubara, B., Liu, S., et al. mobilization and long-term depression. J. Neurosci. 33, 12122–12135. doi: 10.
(2010). Waxholm space: an image-based reference for coordinating mouse 1523/JNEUROSCI.0131-13.2013
brain research. Neuroimage 53, 365–372. doi: 10.1016/j.neuroimage.2010. Oh, S. W., Harris, J. A., Ng, L., Winslow, B., Cain, N., Mihalas, S., et al. (2014).
06.067 A mesoscale connectome of the mouse brain. Nature 508, 207–214. doi: 10.
Junyent, F., and Kremer, E. J. (2015). CAV-2–why a canine virus is a 1038/nature13186
neurobiologist’s best friend. Curr. Opin. Pharmacol. 24, 86–93. doi: 10.1016/j. Palay, S. L. (1956). Synapses in the central nervous system. J. Biophys. Biochem.
coph.2015.08.004 Cytol. 2, 193–202. doi: 10.1083/jcb.2.4.193
Karuppudurai, T., Lin, T.-Y., Ting, C.-Y., Pursley, R., Melnattur, K. V., Diao, F., Palay, S. L., and Palade, G. E. (1955). The fine structure of neurons. J. Biophys.
et al. (2014). A hard-wired glutamatergic circuit pools and relays uv signals Biochem. Cytol. 1, 69–88. doi: 10.1083/jcb.1.1.69
to mediate spectral preference in Drosophila. Neuron 81, 603–615. doi: 10. Park, J., Knezevich, P. L., Wung, W., O’Hanlon, S. N., Goyal, A., Benedetti, K. L.,
1016/j.neuron.2013.12.010 et al. (2011). A conserved juxtacrine signal regulates synaptic partner
Ke, M.-T., Fujimoto, S., and Imai, T. (2013). SeeDB: a simple and morphology- recognition in Caenorhabditis elegans. Neural Dev. 6:28. doi: 10.1186/1749-
preserving optical clearing agent for neuronal circuit reconstruction. Nat. 8104-6-28
Neurosci. 16, 1154–1161. doi: 10.1038/nn.3447 Pech, U., Pooryasin, A., Birman, S., and Fiala, A. (2013). Localization of the
Keller, P. J., Schmidt, A. D., Wittbrodt, J., and Stelzer, E. H. K. (2008). contacts between Kenyon cells and aminergic neurons in the Drosophila
Reconstruction of zebrafish early embryonic development by scanned light melanogaster brain using SplitGFP reconstitution. J. Comp. Neurol. 521,
sheet microscopy. Science 322, 1065–1069. doi: 10.1126/science.1162493 3992–4026. doi: 10.1002/cne.23388
Kim, J., Zhao, T., Petralia, R. S., Yu, Y., Peng, H., Myers, E., et al. (2012). mGRASP Pédelacq, J.-D., Cabantous, S., Tran, T., Terwilliger, T. C., and Waldo, G. S. (2006).
enables mapping mammalian synaptic connectivity with light microscopy. Nat. Engineering and characterization of a superfolder green fluorescent protein.
Methods 9, 96–102. doi: 10.1038/nmeth.1784 Nat. Biotechnol. 24, 79–88. doi: 10.1038/nbt1172
Kopec, C. D., Li, B., Wei, W., Boehm, J., and Malinow, R. (2006). Glutamate Pelkey, K. A., Yuan, X., Lavezzari, G., Roche, K. W., and McBain, C. J. (2007).
receptor exocytosis and spine enlargement during chemically induced long- mGluR7 undergoes rapid internalization in response to activation by the
term potentiation. J. Neurosci. 26, 2000–2009. doi: 10.1523/JNEUROSCI.3918- allosteric agonist AMN082. Neuropharmacology 52, 108–117. doi: 10.1016/j.
05.2006 neuropharm.2006.07.020
Krueger, D. D., Tuffy, L. P., Papadopoulos, T., and Brose, N. (2012). The role Renier, N., Wu, Z., Simon, D. J., Yang, J., Ariel, P., and Tessier-Lavigne, M.
of neurexins and neuroligins in the formation, maturation and function of (2014). iDISCO: a simple, rapid method to immunolabel large tissue samples
vertebrate synapses. Curr. Opin. Neurobiol. 22, 412–422. doi: 10.1016/j.conb. for volume imaging. Cell 159, 896–910. doi: 10.1016/j.cell.2014.10.010
2012.02.012 Richardson, D. S., and Lichtman, J. W. (2015). Clarifying tissue clearing. Cell 162,
Li, Y., Guo, A., and Li, H. (2016). CRASP: CFP reconstitution across synaptic 246–257. doi: 10.1016/j.cell.2015.06.067
partners. Biochem. Biophys. Res. Commun. 469, 352–356. doi: 10.1016/j.bbrc. Russo, S. J., Dietz, D. M., Dumitriu, D., Morrison, J. H., Malenka, R. C., and
2015.12.011 Nestler, E. J. (2010). The addicted synapse: mechanisms of synaptic and
structural plasticity in nucleus accumbens. Trends Neurosci. 33, 267–276. Tsien, J. Z., Huerta, P. T., and Tonegawa, S. (1996). The essential role of
doi: 10.1016/j.tins.2010.02.002 hippocampal CA1 NMDA receptor-dependent synaptic plasticity in spatial
Sankaranarayanan, S., and Ryan, T. A. (2000). Real-time measurements of vesicle- memory. Cell 87, 1327–1338. doi: 10.1016/s0092-8674(00)81827-9
SNARE recycling in synapses of the central nervous system. Nat. Cell Biol. 2, Uttamapinant, C., Tangpeerachaikul, A., Grecian, S., Clarke, S., Singh, U.,
197–204. doi: 10.1038/35008615 Slade, P., et al. (2012). Fast, cell-compatible click chemistry with copper-
Schwarz, L. A., Miyamichi, K., Gao, X. J., Beier, K. T., Weissbourd, B., chelating azides for biomolecular labeling. Angew. Chem. Int. Ed Engl. 51,
DeLoach, K. E., et al. (2015). Viral-genetic tracing of the input-output 5852–5856. doi: 10.1002/anie.201108181
organization of a central noradrenaline circuit. Nature 524, 88–92. doi: 10. Uttamapinant, C., White, K. A., Baruah, H., Thompson, S., Fernández-Suárez, M.,
1038/nature14600 Puthenveetil, S., et al. (2010). A fluorophore ligase for site-specific protein
Shaner, N. C., Lin, M. Z., McKeown, M. R., Steinbach, P. A., Hazelwood, K. L., labeling inside living cells. Proc. Natl. Acad. Sci. U S A 107, 10914–10919.
Davidson, M. W., et al. (2008). Improving the photostability of bright doi: 10.1073/pnas.0914067107
monomeric orange and red fluorescent proteins. Nat. Methods 5, 545–551. Voglmaier, S. M., Kam, K., Yang, H., Fortin, D. L., Hua, Z., Nicoll, R. A., et al.
doi: 10.1038/nmeth.1209 (2006). Distinct endocytic pathways control the rate and extent of synaptic
Shen, Y., Rosendale, M., Campbell, R. E., and Perrais, D. (2014). pHuji, a pH- vesicle protein recycling. Neuron 51, 71–84. doi: 10.1016/j.neuron.2006.
sensitive red fluorescent protein for imaging of exo- and endocytosis. J. Cell 05.027
Biol. 207, 419–432. doi: 10.1083/jcb.201404107 Wang, H., Yang, H., Shivalila, C. S., Dawlaty, M. M., Cheng, A. W., Zhang, F., et al.
Shen, K., Teruel, M. N., Connor, J. H., Shenolikar, S., and Meyer, T. (2013). One-step generation of mice carrying mutations in multiple genes by
(2000). Molecular memory by reversible translocation of calcium/calmodulin- CRISPR/Cas-mediated genome engineering. Cell 153, 910–918. doi: 10.1016/j.
dependent protein kinase II. Nat. Neurosci. 3, 881–886. doi: 10.1038/78783 cell.2013.04.025
Shu, X., Lev-Ram, V., Deerinck, T. J., Qi, Y., Ramko, E. B., Davidson, M. W., Yang, B., Treweek, J. B., Kulkarni, R. P., Deverman, B. E., Chen, C.-K.,
et al. (2011). A genetically encoded tag for correlated light and electron Lubeck, E., et al. (2014). Single-cell phenotyping within transparent intact
microscopy of intact cells, tissues and organisms. PLoS Biol. 9:e1001041. doi: 10. tissue through whole-body clearing. Cell 158, 945–958. doi: 10.1016/j.cell.2014.
1371/journal.pbio.1001041 07.017
Susaki, E. A., Tainaka, K., Perrin, D., Kishino, F., Tawara, T., Watanabe, T. M., Yuan, Q., Xiang, Y., Yan, Z., Han, C., Jan, L. Y., and Jan, Y. N. (2011). Light-
et al. (2014). Whole-brain imaging with single-cell resolution using chemical induced structural and functional plasticity in Drosophila larval visual system.
cocktails and computational analysis. Cell 157, 726–739. doi: 10.1016/j.cell. Science 333, 1458–1462. doi: 10.1126/science.1207121
2014.03.042 Zamanillo, D., Sprengel, R., Hvalby, O., Jensen, V., Burnashev, N., Rozov, A., et al.
Takemoto, K., Matsuda, T., McDougall, M., Klaubert, D. H., Hasegawa, A., (1999). Importance of AMPA receptors for hippocampal synaptic plasticity but
Los, G. V., et al. (2011). Chromophore-assisted light inactivation of HaloTag not for spatial learning. Science 284, 1805–1811. doi: 10.1126/science.284.5421.
fusion proteins labeled with eosin in living cells. ACS Chem. Biol. 6, 401–406. 1805
doi: 10.1021/cb100431e Zhu, Y., Xu, J., and Heinemann, S. F. (2009). Two pathways of synaptic vesicle
Tomer, R., Khairy, K., Amat, F., and Keller, P. J. (2012). Quantitative high-speed retrieval revealed by single-vesicle imaging. Neuron 61, 397–411. doi: 10.1016/j.
imaging of entire developing embryos with simultaneous multiview light-sheet neuron.2008.12.024
microscopy. Nat. Methods 9, 755–763. doi: 10.1038/nmeth.2062
Tomer, R., Ye, L., Hsueh, B., and Deisseroth, K. (2014). Advanced CLARITY for Conflict of Interest Statement: The authors declare that the research was
rapid and high-resolution imaging of intact tissues. Nat. Protoc. 9, 1682–1697. conducted in the absence of any commercial or financial relationships that could
doi: 10.1038/nprot.2014.123 be construed as a potential conflict of interest.
Tour, O., Meijer, R. M., Zacharias, D. A., Adams, S. R., and Tsien, R. Y. (2003).
Genetically targeted chromophore-assisted light inactivation. Nat. Biotechnol. Copyright © 2016 Lee, Oh, Seong and Kim. This is an open-access article distributed
21, 1505–1508. doi: 10.1038/nbt914 under the terms of the Creative Commons Attribution License (CC BY). The use,
Tsetsenis, T., Boucard, A. A., Araç, D., Brunger, A. T., and Südhof, T. C. (2014). distribution and reproduction in other forums is permitted, provided the original
Direct visualization of trans-synaptic neurexin-neuroligin interactions during author(s) or licensor are credited and that the original publication in this journal
synapse formation. J. Neurosci. 34, 15083–15096. doi: 10.1523/JNEUROSCI. is cited, in accordance with accepted academic practice. No use, distribution or
0348-14.2014 reproduction is permitted which does not comply with these terms.
Department of Neuroscience, Mortimer B. Zuckerman Mind Brain Behavior Institute, Kavli Institute for Brain Science,
Columbia University Medical Center, New York, NY, USA
Calcium (Ca2+ ) plays innumerable critical functions in neurons ranging from regulation
of neurotransmitter release and synaptic plasticity to activity-dependent transcription.
Therefore, more than any other cell types, neurons are critically dependent on spatially
and temporally controlled Ca2+ dynamics. This is achieved through an exquisite
level of compartmentalization of Ca2+ storage and release from various organelles.
The function of these organelles in the regulation of Ca2+ dynamics has been
studied for decades using electrophysiological and optical methods combined with
pharmacological and genetic alterations. Mitochondria and the endoplasmic reticulum
(ER) are among the organelles playing the most critical roles in Ca2+ dynamics in
neurons. At presynaptic boutons, Ca2+ triggers neurotransmitter release and synaptic
plasticity, and postsynaptically, Ca2+ mobilization mediates long-term synaptic plasticity.
To explore Ca2+ dynamics in live cells and intact animals, various synthetic and
genetically encoded fluorescent Ca2+ sensors were developed, and recently, many
groups actively increased the sensitivity and diversity of genetically encoded Ca2+
Edited by: indicators (GECIs). Following conjugation with various signal peptides, these improved
Marc Fivaz, GECIs can be targeted to specific subcellular compartments, allowing monitoring
Duke NUS Graduate Medical School,
Singapore of organelle-specific Ca2+ dynamics. Here, we review recent findings unraveling
Reviewed by: novel roles for mitochondria- and ER-dependent Ca2+ dynamics in neurons and
Lucas Pozzo-Miller, at synapses.
University of Alabama at Birmingham,
USA Keywords: synapse, mitochondria, endoplasmic reticulum, circuit function, calcium dynamics
William N. Green,
University of Chicago, USA
*Correspondence:
INTRODUCTION
Franck Polleux
[email protected] Calcium (Ca2+ ) ions govern prevalent physiological processes in various cell types (Rizzuto
and Pozzan, 2006; Clapham, 2007). This is especially prominent in excitable cells like neurons
Received: 13 June 2016 where Ca2+ influx through the plasma membrane and release of Ca2+ from internal stores
Accepted: 30 August 2016 transduce the effects of changes in membrane polarization and therefore mediate faithful
Published: 16 September 2016 transfer or storage of information over various timescales (milliseconds to minutes/hours).
Citation: Therefore, regulation of intracellular Ca2+ homeostasis is central to the proper function of
Kwon S-K, Hirabayashi Y and neuronal circuits. The maintenance of baseline levels of intracellular Ca2+ levels is regulated
Polleux F (2016) Organelle-Specific
in part through exchangers and pumps such as the plasma membrane Ca2+ -ATPase (PMCA
Sensors for Monitoring Ca2+
Dynamics in Neurons.
pump), the Na+ /Ca2+ exchanger (NCX), and the Na+ /Ca2+ -K+ exchanger (NCKX) which
Front. Synaptic Neurosci. 8:29. extrude Ca2+ through the plasma membrane into the extracellular space. In addition to
doi: 10.3389/fnsyn.2016.00029 these mechanisms, intracellular organelles, such as mitochondria and endoplasmic reticulum
(ER), are able to regulate cytoplasmic Ca2+ ([Ca2+ ]c ) [Ca2+ ]c . This has been characterized in various species,
through mitochondrial calcium uniporter (MCU) and smooth neuronal cell types and circuits (Figure 1A). At the Drosophila
endoplasmic reticulum Ca2+ -ATPase (SERCA), respectively. neuromuscular junction (NMJ), the GTPase dMiro mutant lacks
In neurons, mitochondria and ER play important presynaptic mitochondria through impaired axonal transport
physiological roles via [Ca2+ ]c regulation, thereby diverse (Guo et al., 2005; Wang and Schwarz, 2009). During prolonged
synaptic functions including basal synaptic transmission, stimulation, these mutants lacking presynaptic mitochondria
presynaptic short-term plasticity, and long-term plasticity can displayed subtle, but significantly increased presynaptic Ca2+
be regulated by these organelles (Verkhratsky, 2005; Bardo et al., accumulation and display decrease forms of sustained synaptic
2006; Mattson et al., 2008; Vos et al., 2010). In addition, impaired transmission or synaptic ‘‘fatigue’’ (Guo et al., 2005). Drosophila
Ca2+ homeostasis in the nervous system has been proposed Drp1 mutants also deplete presynaptic mitochondria at NMJ
to play an important function in the physio-pathological and exhibit elevated presynaptic Ca2+ levels in resting and
mechanisms underlying Alzheimer’s disease, Parkinson’s evoked states. However, spontaneous release (mini Excitatory
disease, and spinocerebellar ataxia (Verkhratsky, 2005; Mattson junctional potential, mEJP) was not altered, but the evoked
et al., 2008; Schon and Przedborski, 2011). synaptic transmission was impaired during high frequency
To monitor Ca2+ dynamics, various fluorescent Ca2+ dyes stimulation, and this defect was partially rescued by ATP
and genetically encoded Ca2+ indicators (GECIs) were developed (Verstreken et al., 2005) suggesting that mitochondria plays a role
and applied both in vitro and in vivo. Also, GECIs tagged with in synaptic transmission through their ability to generate ATP
target peptide sequences have allowed imaging of Ca2+ dynamics through oxidative phosphorylation. Although mitochondrial
in specific organelles (Rizzuto et al., 1992; Palmer et al., 2004; Ca2+ uptake has limited effects on Drosophila NMJ neurons, in
Palmer and Tsien, 2006). mammalian NMJ terminals, acute inhibition of mitochondrial
Previously published reviews have already summarized the Ca2+ uptake causes rapid depression of the endplate potential
usefulness and limitations of various Ca2+ sensors and GECIs (EPP) and increased asynchronous release (David and Barrett,
applied to neuronal and non-neuronal cells (Palmer and Tsien, 2003). Furthermore, in synapses of the mammalian central
2006; Knopfel, 2012; Tian et al., 2012; Rose et al., 2014). In this nervous system (CNS), mitochondria-dependent Ca2+ uptake
review, we only describe recently uncovered insights about Ca2+ accelerates the recovery from synaptic depression in the calyx of
dynamics and its regulation by mitochondria and ER, and we Held (Billups and Forsythe, 2002). Other studies in mammalian
discuss how these organelle-specific Ca2+ sensors have been used hippocampal neurons claimed that impaired mitochondrial
for the exploration of the role of these subcellular compartments anchoring at presynaptic sites increases presynaptic Ca2+ during
in the regulation of Ca2+ homeostasis and synaptic function in repetitive stimulation and produces short-term facilitation (STF),
neurons. and insulin-like growth factor-1 receptor (IGF-1R) signaling
regulates resting mitochondrial Ca2+ level and spontaneous
UNVEILED SYNAPTIC FUNCTIONS OF transmission (Kang et al., 2008; Gazit et al., 2016). Although
MITOCHONDRIA-DEPENDENT Ca2+ most pharmacological studies employed uncoupling agents as
HOMEOSTASIS mitochondrial Ca2+ influx blocker, which may affect ATP
production, these reports support presynaptic control via
Mitochondrial Ca2+ uptake has been studied since the 1950s mitochondrial Ca2+ import (Ly and Verstreken, 2006). A recent
from studies of rat heart muscle and kidney (Slater and study demonstrates that presynaptic boutons associated with
Cleland, 1953; Deluca and Engstrom, 1961). In the nervous mitochondria display lower levels of [Ca2+ ]c accumulation than
system, mitochondria were described at presynaptic terminals presynaptic boutons not associated with mitochondria (Kwon
and dendrites of various neuronal subtypes using the light and et al., 2016). Furthermore, acute inhibition of mitochondria
electron microscope (EM) several decades ago (Bartelmez and calcium import increased [Ca2+ ]c accumulation at presynaptic
Hoerr, 1933; Palay, 1956; Gray, 1963; Shepherd and Harris, 1998; boutons occupied by mitochondria. In the same study, we
Rowland et al., 2000). In axons, mitochondria are short and demonstrate that this mitochondria-dependent regulation of
sparsely distributed, and interestingly, several studies showed [Ca2+ ]c plays an important role in regulating presynaptic release
that half of presynaptic boutons are occupied by mitochondria properties including spontaneous release, asynchronous release
(Shepherd and Harris, 1998; Kang et al., 2008). In contrast, and short-term synaptic plasticity.
dendritic mitochondria have tubular shapes and they are rarely In addition to regulation of [Ca2+ ]c clearance, Ca2+ release
observed in postsynaptic spines in the excitatory neurons (Sheng from mitochondria plays important roles at presynaptic
and Hoogenraad, 2007; Kasthuri et al., 2015). sites (Figure 1A). Following the sustained high frequency
At presynaptic boutons and terminals, synaptic vesicle stimulation, an enhancement of synaptic transmission
(SV) fusion with the plasma membrane occurs following lasting tens of seconds to minutes is observed and which
increase of [Ca2+ ]c following opening of voltage-sensitive Ca2+ is called post-tetanic potentiation (PTP; Zucker, 1989).
channels (VSCC) followed by Ca2+ binding to sensors like Mitochondrial Ca2+ release is suggested as one of the underlying
synaptotagmins (Schneggenburger and Neher, 2005; Neher and mechanisms for this prolonged enhancement of synaptic
Sakaba, 2008; Jahn and Fasshauer, 2012; Südhof, 2012). The transmission. Pharmacological inhibition of mitochondrial
ability of mitochondria to import Ca2+ into the mitochondrial Ca2+ uptake and release at the crayfish NMJ impaired PTP
matrix ([Ca2+ ]m ) plays a role in regulating presynaptic (Tang and Zucker, 1997; Zhong et al., 2001). Furthermore,
FIGURE 1 | Synaptic functions regulated by endoplasmic reticulum (ER) and mitochondria-dependent Ca2+ homeostasis. (A) Schematic diagram
depicting the presynaptic functions regulated by ER- and mitochondria-dependent Ca2+ dynamics. Ca2+ release from ER can modulate spontaneous
neurotransmitter release, short-term facilitation (STF) and long-term depression (LTD). Ca2+ re-uptake by the ER controls the spontaneous release and STF.
Presynaptic mitochondria also play important roles in regulating spontaneous neurotransmitter release, STF and post-tetanic potentiation (PTP) through their ability to
regulate Ca2+ clearance. (B) A simplified schematic diagram depicting the postsynaptic functions regulated by ER- and mitochondria-dependent Ca2+ dynamics.
Ca2+ release from ER via IP3 -induced Ca2+ release (IICR) and Ca2+ -induced Ca2+ release (CICR) controls long-term potentiation (LTP) and LTD. In fact, depending
on neuronal and synaptic subtypes, IP3 R and RyR show differential distribution and distinct synaptic functions. Dendritic mitochondrial Ca2+ influx can regulate ATP
synthesis, Ca2+ homeostasis and dendritic development. In non-neuronal cell types, direct Ca2+ exchange between ER and mitochondria have been described, but
their role in neurons has not yet been documented. IP3 R, IP3 receptor; RyR, ryanodine receptor; SERCA, smooth endoplasmic reticulum Ca2+ -ATPase; VGCC,
voltage-gated Ca2+ channel; PMCA, plasma membrane Ca2+ -ATPase; NCX: the Na+ /Ca2+ exchanger; mPTP, mitochondrial permeability transition pore; MCU,
mitochondrial calcium uniporter; VDAC, voltage-dependent anion channel; mGluR, metabotropic glutamate receptor; GluN, NMDA receptor.
similar phenotypes were observed at mouse NMJ and contain mitochondria (Chicurel and Harris, 1992). However, a
hippocampal mossy fiber synapses with blocking the physiological role of these postsynaptic mitochondria is largely
mitochondrial NCX, which mediates mitochondrial Ca2+ unknown. In general, mitochondria are distributed primarily in
release (García-Chacón et al., 2006; Lee et al., 2007). dendrite shaft and therefore localized microns away from the
In contrast to presynaptic boutons and terminals, the postsynaptic density, but might still be able to buffer [Ca2+ ]c
postsynaptic function of mitochondrial Ca2+ regulation is less mobilized through Ca2+ -channels and glutamate receptors
well-documented. In mouse hippocampal pyramidal neurons (Thayer and Miller, 1990; White and Reynolds, 1995;
(Li et al., 2004), a minority (<5%) of dendritic spines contains Wang and Thayer, 2002). This mitochondrial calcium import
mitochondria. Also, large branched spines in hippocampal CA3 can stimulate tricarboxylic acid (TCA) cycle and might
increase ATP production (Kann and Kovács, 2007) and Ca2+ (Gazit et al., 2016; Kwon et al., 2016; Marland et al.,
may also regulate other ATP-dependent Ca2+ pumps like 2016). Both sensors displayed action potential (AP)-dependent
PMCA and SERCA. While it is still unclear whether or not mitochondrial Ca2+ import. In addition, red fluorescent GECIs
mitochondria play significant roles in regulating postsynaptic by replacing cpEGFP with cpmApple or cpmRuby (mtRCaMP1e
[Ca2+ ]c under physiological conditions of neurotransmission, and LAR-GECO1.2) revealed mitochondrial Ca2+ import
they might play a role in pathophysiological contexts. For simultaneously with cytosolic Ca2+ (Akerboom et al., 2013; Wu
example, neurons lacking LRRK2, a protein associated with et al., 2014).
Parkinson’s disease, show impaired dendritic Ca2+ homeostasis However, these fluorescent proteins have some limitations,
through mitochondrial defects and thought to cause defective for example, they are affected by pH and mitochondrial matrix
mitochondrial depolarization and reduction in dendritic pH (pHm ) can be changed by Ca2+ influx (Abad et al., 2004;
complexity (Figure 1B; Cherra et al., 2013). Poburko et al., 2011; Chouhan et al., 2012; Marland et al.,
Overall, mitochondria-dependent Ca2+ clearance and release 2016). In addition to this point, [Ca2+ ]m can span broad ranges
in neurons plays important physiological and developmental (0.05–300 µM) depending on cell types and stimulation protocol
roles pre- and post-synaptically but their functional importance (Arnaudeau et al., 2001; Palmer and Tsien, 2006). Thus, Kd
seems to depend on the neuronal subtypes and the structure/size value for Ca2+ of mitochondrial GECI should be considered for
of the pre- and postsynaptic compartments. experimental purposes because high affinity (low Kd ) sensors can
be easily saturated by high [Ca2+ ]m and low affinity (high Kd )
MITOCHONDRIAL Ca2+ -IMAGING IN sensors may not be sensitive enough to detect small [Ca2+ ]m
NEURONS AND AT SYNAPSES changes. Several studies reported low affinity mitochondrial
Ca2+ probes for avoiding saturation (Arnaudeau et al., 2001;
To investigate organelle-specific Ca2+ dynamics, various Ca2+ Suzuki et al., 2014).
sensors are developed (Table 1). One of the first method In conclusion, these mitochondria-targeted GECIs allow
developed to monitor mitochondrial Ca2+ dynamics was imaging of mitochondria Ca2+ dynamics in neurons and have
established using rhod-2, a cationic chemical Ca2+ -binding revealed interesting, synapse-specific properties of mitochondria
fluorophore preferentially accumulating in the mitochondrial in the regulation of [Ca2+ ]c and neurotransmitter release
matrix presumably because of the highly negative membrane properties.
potential across the mitochondrial inner membrane (Minta
et al., 1989). Then, in the calyx of Held, rhod-2 and rhod-
FF (low affinity version) were used to visualize presynaptic REGULATION OF SYNAPTIC Ca2+
mitochondrial Ca2+ transient (Billups and Forsythe, 2002). DYNAMICS BY THE ENDOPLASMIC
However, these dyes cannot be precisely targeted to these RETICULUM
organelles. Therefore, GECIs have recently become the
preferred method to image Ca2+ in specific organelles including Neurons are among the most polarized cell types in our body
mitochondria. For mitochondrial matrix localization, the and consists of a soma, relatively short dendrites and long
targeting presequence of subunit VIII of human cytochrome c axons. ER is found throughout the entire length of neuronal
oxidase (COXVIII) was tagged to GECIs (Rizzuto et al., 1992). processes, and usually rough ER is prominent in the cell
Mitochondria-targeted aequorin (mt-AEQ), a luminescent body and proximal dendrites, whereas smooth ER is dominant
Ca2+ indicator, was first employed to monitor the neuronal in distal dendrites, spines and axons (Spacek and Harris,
mitochondrial Ca2+ , and this probe showed NMDA-induced 1997; Verkhratsky, 2005). ER imports and sequesters large
mitochondrial Ca2+ increase in hippocampal neurons (Baron amount of Ca2+ ([Ca2+ ]er ∼500 µM) through SERCA and
et al., 2003). However, this probe needs a chemical reaction store-operated Ca2+ entry (SOCE) mechanism (Verkhratsky,
characterized by a modest turnover rate and has very limited 2005; Bardo et al., 2006). Ca2+ release from ER is mediated
dynamic range (Palmer and Tsien, 2006). Other GECIs have by two major mechanisms, called Ca2+ -induced Ca2+ release
been developed and tested in various neuronal subtypes (CICR) and IP3 -induced Ca2+ release (IICR; Verkhratsky, 2005;
with the same targeting sequence. Mitochondrial-targeted Bardo et al., 2006). CICR is caused by the cytosolic Ca2+
ratiometric pericam (2mtRP) consists of circularly permutated increase through N-Methyl-D-Aspartate receptors (NMDAR,
Enhanced yellow fluorescent protein (cpEFYP) conjugated GluN receptors) and voltage-gated Ca2+ channels (VGCCs),
with Ca2+ -responsive calmodulin (CaM) and its binding whereas IICR is triggered by IP3 , which is generated via activation
peptide (Nagai et al., 2001; Robert et al., 2001). This probe of phospholipase C (PLC) depending on metabotropic glutamate
has a bimodal excitation spectrum and the relative emission receptors (mGluRs) or other receptors like receptor tyrosine
intensity is dependent on Ca2+ -binding. In hippocampal kinases (Figure 1).
neurons, the use of 2mtRP described mitochondrial Ca2+ Ryanodine receptors (RyRs) are involved in CICR, and they
uptake and also determined cytosolic Ca2+ rise upon synaptic have three major subtypes; RyR1, RyR2, and RyR3. All of these
activation via dual imaging with cytosolic Ca2+ dye (fura-red isoforms are detected in the brain, and show region-specific
AM; Young et al., 2008). Other CaM conjugated cpEGFPs expression (Sharp et al., 1993; Furuichi et al., 1994; Giannini
called GCaMPs (mito-GCaMP2, 2mtGCaMP6m, and mito- et al., 1995; Verkhratsky, 2005; Bardo et al., 2006; Baker et al.,
GCaMP5G) were used to monitor axonal mitochondrial 2013). Similar to RyRs, IP3 receptor (IP3 R), which mediate IICR,
Mitochondria Dye The calyx of Held 0.57, 19 575 590 3.4 Billups and Forsythe (2002)
Rhod-2, Rhod-FF
GECI
mito-aequorin Hippocampal 1–2 Luminescence Baron et al. (2003)
(Hp) neuron
2mtRP (ratioPericam) Hp neuron 1.7 Ratiometric, 535/20 10 Young et al. (2008)
405/485
mito-GCaMP2 Hp neuron 0.195 488 507 5 Marland et al. (2016)
2mtGCaMP6m Hp neuron 0.167 488 510 38 Patron et al. (2014),
Gazit et al. (2016)
mtRCaMP1e Cortical neuron 1.6 572 592.5 6.5 Akerboom et al. (2013)
LAR-GECO1.2 DRG and Hp 12 561 589 Wu et al. (2014)
neurons
ER Dye
Mag-Fura-2 Sensory neuron 53 Ratiometric, 510 25 Solovyova et al. (2002)
340/380
GECI
D1ER Hp neuron 0.8, 60 FRET, 450 475/40, 1.6 Zhang et al. (2010)
535/25
erGAP1 DRG neuron, Hp 12 Luminescence, 510 3∼4 Rodriguez-Garcia et al. (2014)
slice 403/470
G-CEPIA1er Cerebellar Purkinje 672 488 511 4.7 ± 0.3 Suzuki et al. (2014)
cell
GCaMPer (10.19) Cortical neurons 400 490 540/50 14 Henderson et al. (2015)
consist of three isoforms, IP3 R1, IP3 R2, and IP3 R3, but IP3 R1 et al., 2000; Nishiyama et al., 2000; Raymond and Redman, 2002).
is the dominant form in the brain (Sharp et al., 1993, 1999; Hippocampal mossy fiber pathway (MF, dentate gyrus to CA3)
Verkhratsky, 2005; Bardo et al., 2006; Baker et al., 2013). shows IICR- and CICR-dependent LTP and LTD, however, there
Long-term synaptic plasticity is regulated by Ca2+ - are conflicting results regarding the underlying mechanisms
dependent signaling mechanisms such as Ca2+ /calmodulin- (Figure 1B; Yeckel et al., 1999; Itoh et al., 2001; Kapur et al., 2001;
dependent kinase II (CaMKII), calcineurin (a Ca2+ -dependent Mellor and Nicoll, 2001; Lauri et al., 2003; Lei et al., 2003).
phosphatase), protein phosphatase 1 (PP1) and protein kinase C Presynaptic ER-dependent Ca2+ release is also detected and
(PKC; Malenka and Nicoll, 1999; Yang et al., 1999; Lüscher and contributes to changes in neurotransmitter release properties
Malenka, 2012). Therefore, Ca2+ release from intracellular stores and short-term synaptic plasticity at various inhibitory and
like the ER regulates long-term synaptic plasticity in specific excitatory synapses including basket cell to Purkinje cell
circuits. synapses, hippocampal MF pathway, SC-CA1 and CA3-CA3
In cerebellar Purkinje cell dendrites, mGluR-IP3 -dependent pyramidal neuron synapses (Figure 1A; Llano et al., 2000;
Ca2+ increase is observed during parallel fiber (PF) stimulation Emptage et al., 2001; Liang et al., 2002; Galante and Marty, 2003;
and this mediates long-term depression (LTD) of PF-Purkinje Lauri et al., 2003; Sharma and Vijayaraghavan, 2003; Unni et al.,
cell pathway (Finch and Augustine, 1998; Takechi et al., 1998; 2004; Mathew and Hablitz, 2008).
Miyata et al., 2000; Wang et al., 2000). At synapses made In addition to Ca2+ efflux, Ca2+ uptake by ER via SERCA
by hippocampal Schaffer collateral (SC) onto CA1 pyramidal pump affects STF at SC-CA1 presynapses and NMJ (Figure 1A;
neurons, both long-term potentiation (LTP) and LTD are linked Castonguay and Robitaille, 2001; Scullin and Partridge, 2010;
to IP3 -dependent signaling (Oliet et al., 1997; Nishiyama et al., Scullin et al., 2010). Stromal interaction molecules (STIMs)
2000; Raymond and Redman, 2002; Nagase et al., 2003). In and Orai1, which allow SOCE, are localized to neuronal
addition, CICR is also observed in CA1 pyramidal neuronal compartment including dendritic spines, and impaired SOCE
spines, and LTD is abolished in RyR3-deficient mice and alters α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid
following application of RyR inhibitor (ryanodine) although receptor (AMPAR) trafficking, neuronal Ca2+ signaling and LTP
the connection between CICR and LTP is controversial in SC- in CA1 pyramidal neuron and cerebellar Purkinje neurons (Baba
CA1 pathway (Reyes and Stanton, 1996; Emptage et al., 1999; et al., 2003; Hartmann et al., 2014; Korkotian et al., 2014; Garcia-
Futatsugi et al., 1999; Sandler and Barbara, 1999; Kovalchuk Alvarez et al., 2015; Segal and Korkotian, 2015).
REFERENCES Arnaudeau, S., Kelley, W. L., Walsh, J. V. Jr., and Demaurex, N. (2001).
Mitochondria recycle Ca2+ to the endoplasmic reticulum and prevent the
Abad, M. F., Di Benedetto, G., Magalhães, P. J., Filippin, L., and Pozzan, T. (2004). depletion of neighboring endoplasmic reticulum regions. J. Biol. Chem. 276,
Mitochondrial pH monitored by a new engineered green fluorescent protein 29430–29439. doi: 10.1074/jbc.m103274200
mutant. J. Biol. Chem. 279, 11521–11529. doi: 10.1074/jbc.m306766200 Baba, A., Yasui, T., Fujisawa, S., Yamada, R. X., Yamada, M. K., Nishiyama, N.,
Akerboom, J., Carreras Calderón, N., Tian, L., Wabnig, S., Prigge, M., Tolö, J., et al. et al. (2003). Activity-evoked capacitative Ca2+ entry: implications in synaptic
(2013). Genetically encoded calcium indicators for multi-color neural activity plasticity. J. Neurosci. 23, 7737–7741.
imaging and combination with optogenetics. Front. Mol. Neurosci. 6:2. doi: 10. Baker, K. D., Edwards, T. M., and Rickard, N. S. (2013). The role of
3389/fnmol.2013.00002 intracellular calcium stores in synaptic plasticity and memory consolidation.
Neurosci. Biobehav. Rev. 37, 1211–1239. doi: 10.1016/j.neubiorev.2013. Fujimoto, M., and Hayashi, T. (2011). New insights into the role of mitochondria-
04.011 associated endoplasmic reticulum membrane. Int. Rev. Cell Mol. Biol. 292,
Bardo, S., Cavazzini, M. G., and Emptage, N. (2006). The role of the endoplasmic 73–117. doi: 10.1016/B978-0-12-386033-0.00002-5
reticulum Ca2+ store in the plasticity of central neurons. Trends Pharmacol. Furuichi, T., Furutama, D., Hakamata, Y., Nakai, J., Takeshima, H., and
Sci. 27, 78–84. doi: 10.1016/j.tips.2005.12.008 Mikoshiba, K. (1994). Multiple types of ryanodine receptor/Ca2+ release
Baron, K. T., Wang, G. J., Padua, R. A., Campbell, C., and Thayer, S. A. channels are differentially expressed in rabbit brain. J. Neurosci. 14, 4794–4805.
(2003). NMDA-evoked consumption and recovery of mitochondrially targeted Futatsugi, A., Kato, K., Ogura, H., Li, S. T., Nagata, E., Kuwajima, G., et al. (1999).
aequorin suggests increased Ca2+ uptake by a subset of mitochondria in Facilitation of NMDAR-independent LTP and spatial learning in mutant mice
hippocampal neurons. Brain Res. 993, 124–132. doi: 10.1016/j.brainres.2003. lacking ryanodine receptor type 3. Neuron 24, 701–713. doi: 10.1016/s0896-
09.022 6273(00)81123-x
Bartelmez, G. W., and Hoerr, N. L. (1933). The vestibular club endings in Galante, M., and Marty, A. (2003). Presynaptic ryanodine-sensitive calcium stores
ameiurus. Further evidence on the morphology of the synapse. J. Comp. Neurol. contribute to evoked neurotransmitter release at the basket cell-Purkinje cell
57, 401–428. doi: 10.1002/cne.900570303 synapse. J. Neurosci. 23, 11229–11234.
Billups, B., and Forsythe, I. D. (2002). Presynaptic mitochondrial calcium Garcia-Alvarez, G., Lu, B., Yap, K. A., Wong, L. C., Thevathasan, J. V., Lim, L.,
sequestration influences transmission at mammalian central synapses. et al. (2015). STIM2 regulates PKA-dependent phosphorylation and trafficking
J. Neurosci. 22, 5840–5847. of AMPARs. Mol. Biol. Cell 26, 1141–1159. doi: 10.1091/mbc.e14-07-1222
Cárdenas, C., Miller, R. A., Smith, I., Bui, T., Molgo, J., Muller, M., et al. García-Chacón, L. E., Nguyen, K. T., David, G., and Barrett, E. F. (2006).
(2010). Essential regulation of cell bioenergetics by constitutive InsP3 receptor Extrusion of Ca2+ from mouse motor terminal mitochondria via a Na+ -
Ca2+ transfer to mitochondria. Cell 142, 270–283. doi: 10.1016/j.cell.2010. Ca2+ exchanger increases post-tetanic evoked release. J. Physiol. 574, 663–675.
06.007 doi: 10.1113/jphysiol.2006.110841
Castonguay, A., and Robitaille, R. (2001). Differential regulation of transmitter Gazit, N., Vertkin, I., Shapira, I., Helm, M., Slomowitz, E., Sheiba, M., et al. (2016).
release by presynaptic and glial Ca2+ internal stores at the neuromuscular IGF-1 receptor differentially regulates spontaneous and evoked transmission
synapse. J. Neurosci. 21, 1911–1922. via mitochondria at hippocampal synapses. Neuron 89, 583–597. doi: 10.1016/j.
Cherra, S. J. III, Steer, E., Gusdon, A. M., Kiselyov, K., and Chu, C. T. neuron.2015.12.034
(2013). Mutant LRRK2 elicits calcium imbalance and depletion of dendritic Giannini, G., Conti, A., Mammarella, S., Scrobogna, M., and Sorrentino, V. (1995).
mitochondria in neurons. Am. J. Pathol. 182, 474–484. doi: 10.1016/j.ajpath. The ryanodine receptor/calcium channel genes are widely and differentially
2012.10.027 expressed in murine brain and peripheral tissues. J. Cell Biol. 128, 893–904.
Chicurel, M. E., and Harris, K. M. (1992). Three-dimensional analysis of the doi: 10.1083/jcb.128.5.893
structure and composition of CA3 branched dendritic spines and their synaptic Gray, E. G. (1963). Electron microscopy of presynaptic organelles of the spinal
relationships with mossy fiber boutons in the rat hippocampus. J. Comp. cord. J. Anat. 97, 101–106.
Neurol. 325, 169–182. doi: 10.1002/cne.903250204 Guo, X., Macleod, G. T., Wellington, A., Hu, F., Panchumarthi, S., Schoenfield, M.,
Chouhan, A. K., Ivannikov, M. V., Lu, Z., Sugimori, M., Llinas, R. R., and et al. (2005). The GTPase dMiro is required for axonal transport of
Macleod, G. T. (2012). Cytosolic calcium coordinates mitochondrial energy mitochondria to Drosophila synapses. Neuron 47, 379–393. doi: 10.1016/j.
metabolism with presynaptic activity. J. Neurosci. 32, 1233–1243. doi: 10. neuron.2005.06.027
1523/jneurosci.1301-11.2012 Hartmann, J., Karl, R. M., Alexander, R. P., Adelsberger, H., Brill, M. S.,
Clapham, D. E. (2007). Calcium signaling. Cell 131, 1047–1058. doi: 10. Rühlmann, C., et al. (2014). STIM1 controls neuronal Ca2+ signaling, mGluR1-
1016/j.cell.2007.11.028 dependent synaptic transmission and cerebellar motor behavior. Neuron 82,
Csordás, G., Golenár, T., Seifert, E. L., Kamer, K. J., Sancak, Y., Perocchi, F., et al. 635–644. doi: 10.1016/j.neuron.2014.03.027
(2013). MICU1 controls both the threshold and cooperative activation of the Henderson, M. J., Baldwin, H. A., Werley, C. A., Boccardo, S., Whitaker, L. R.,
mitochondrial Ca2+ uniporter. Cell Metab. 17, 976–987. doi: 10.1016/j.cmet. Yan, X., et al. (2015). A low affinity GCaMP3 variant (GCaMPer) for imaging
2013.04.020 the endoplasmic reticulum calcium store. PLoS One 10:e0139273. doi: 10.
Csordás, G., Várnai, P., Golenár, T., Roy, S., Purkins, G., Schneider, T. G., et al. 1371/journal.pone.0139273
(2010). Imaging interorganelle contacts and local calcium dynamics at the ER- Itoh, S., Ito, K., Fujii, S., Kaneko, K., Kato, K., Mikoshiba, K., et al. (2001). Neuronal
mitochondrial interface. Mol. Cell 39, 121–132. doi: 10.1016/j.molcel.2010.06. plasticity in hippocampal mossy fiber-CA3 synapses of mice lacking the
029 inositol-1,4,5-trisphosphate type 1 receptor. Brain Res. 901, 237–246. doi: 10.
David, G., and Barrett, E. F. (2003). Mitochondrial Ca2+ uptake prevents 1016/s0006-8993(01)02373-3
desynchronization of quantal release and minimizes depletion during repetitive Jahn, R., and Fasshauer, D. (2012). Molecular machines governing exocytosis of
stimulation of mouse motor nerve terminals. J. Physiol. 548, 425–438. doi: 10. synaptic vesicles. Nature 490, 201–207. doi: 10.1038/nature11320
1113/jphysiol.2002.035196 Kang, J. S., Tian, J. H., Pan, P. Y., Zald, P., Li, C., Deng, C., et al. (2008). Docking of
De Stefani, D., Patron, M., and Rizzuto, R. (2015). Structure and function of axonal mitochondria by syntaphilin controls their mobility and affects short-
the mitochondrial calcium uniporter complex. Biochim. Biophys. Acta 1853, term facilitation. Cell 132, 137–148. doi: 10.1016/j.cell.2007.11.024
2006–2011. doi: 10.1016/j.bbamcr.2015.04.008 Kann, O., and Kovács, R. (2007). Mitochondria and neuronal activity. Am. J.
Deluca, H. F., and Engstrom, G. W. (1961). Calcium uptake by rat kidney Physiol. Cell Physiol. 292, C641–C657. doi: 10.1152/ajpcell.00222.2006
mitochondria. Proc. Natl. Acad. Sci. U S A 47, 1744–1750. doi: 10.1073/pnas. Kapur, A., Yeckel, M., and Johnston, D. (2001). Hippocampal mossy fiber
47.11.1744 activity evokes Ca2+ release in CA3 pyramidal neurons via a metabotropic
Emptage, N., Bliss, T. V., and Fine, A. (1999). Single synaptic events evoke NMDA glutamate receptor pathway. Neuroscience 107, 59–69. doi: 10.1016/s0306-
receptor-mediated release of calcium from internal stores in hippocampal 4522(01)00293-7
dendritic spines. Neuron 22, 115–124. doi: 10.1016/s0896-6273(00)80683-2 Kasthuri, N., Hayworth, K. J., Berger, D. R., Schalek, R. L., Conchello, J. A.,
Emptage, N. J., Reid, C. A., and Fine, A. (2001). Calcium stores in hippocampal Knowles-Barley, S., et al. (2015). Saturated reconstruction of a volume of
synaptic boutons mediate short-term plasticity, store-operated Ca2+ entry neocortex. Cell 162, 648–661. doi: 10.1016/j.cell.2015.06.054
and spontaneous transmitter release. Neuron 29, 197–208. doi: 10.1016/s0896- Knopfel, T. (2012). Genetically encoded optical indicators for the analysis of
6273(01)00190-8 neuronal circuits. Nat. Rev. Neurosci. 13, 687–700. doi: 10.1523/JNEUROSCI.
Fieni, F., Lee, S. B., Jan, Y. N., and Kirichok, Y. (2012). Activity of the 0381-14.2014
mitochondrial calcium uniporter varies greatly between tissues. Nat. Commun. Korkotian, E., Frotscher, M., and Segal, M. (2014). Synaptopodin regulates spine
3:1317. doi: 10.1038/ncomms2325 plasticity: mediation by calcium stores. J. Neurosci. 34, 11641–11651. doi: 10.
Finch, E. A., and Augustine, G. J. (1998). Local calcium signalling by inositol- 1523/JNEUROSCI.0381-14.2014
1,4,5-trisphosphate in Purkinje cell dendrites. Nature 396, 753–756. doi: 10. Kornmann, B. (2013). The molecular hug between the ER and the mitochondria.
1038/25541 Curr. Opin. Cell Biol. 25, 443–448. doi: 10.1016/j.ceb.2013.02.010
Kovalchuk, Y., Eilers, J., Lisman, J., and Konnerth, A. (2000). NMDA receptor- Neher, E., and Sakaba, T. (2008). Multiple roles of calcium ions in the regulation
mediated subthreshold Ca2+ signals in spines of hippocampal neurons. of neurotransmitter release. Neuron 59, 861–872. doi: 10.1016/j.neuron.2008.
J. Neurosci. 20, 1791–1799. 08.019
Kwon, S., Sando, R. III, Lewis, T. L., Hirabayashi, Y., Maximov, A., and Polleux, F. Nishiyama, M., Hong, K., Mikoshiba, K., Poo, M. M., and Kato, K. (2000). Calcium
(2016). LKB1 regulates mitochondria-dependent presynaptic calcium stores regulate the polarity and input specificity of synaptic modification.
clearance and neurotransmitter release properties at excitatory synapses Nature 408, 584–588. doi: 10.1038/35046067
along cortical axons. PLoS Biol. 14:e1002516. doi: 10.1371/journal.pbio. Okubo, Y., Suzuki, J., Kanemaru, K., Nakamura, N., Shibata, T., and Iino, M.
1002516 (2015). Visualization of Ca2+ filling mechanisms upon synaptic inputs
Lauri, S. E., Bortolotto, Z. A., Nistico, R., Bleakman, D., Ornstein, P. L., Lodge, D., in the endoplasmic reticulum of cerebellar purkinje cells. J. Neurosci. 35,
et al. (2003). A role for Ca2+ stores in kainate receptor-dependent synaptic 15837–15846. doi: 10.1523/JNEUROSCI.3487-15.2015
facilitation and LTP at mossy fiber synapses in the hippocampus. Neuron 39, Oliet, S. H., Malenka, R. C., and Nicoll, R. A. (1997). Two distinct forms of
327–341. doi: 10.1016/s0896-6273(03)00369-6 long-term depression coexist in CA1 hippocampal pyramidal cells. Neuron 18,
Lee, D., Lee, K. H., Ho, W. K., and Lee, S. H. (2007). Target cell-specific 969–982. doi: 10.1016/s0896-6273(00)80336-0
involvement of presynaptic mitochondria in post-tetanic potentiation at Palay, S. L. (1956). Synapses in the central nervous system. J. Biophys. Biochem.
hippocampal mossy fiber synapses. J. Neurosci. 27, 13603–13613. doi: 10. Cytol. 2, 193–202. doi: 10.1083/jcb.2.4.193
1523/JNEUROSCI.3985-07.2007 Palmer, A. E., Jin, C., Reed, J. C., and Tsien, R. Y. (2004). Bcl-2-mediated
Lei, S., Pelkey, K. A., Topolnik, L., Congar, P., Lacaille, J. C., and McBain, C. J. alterations in endoplasmic reticulum Ca2+ analyzed with an improved
(2003). Depolarization-induced long-term depression at hippocampal mossy genetically encoded fluorescent sensor. Proc. Natl. Acad. Sci. U S A 101,
fiber-CA3 pyramidal neuron synapses. J. Neurosci. 23, 9786–9795. 17404–17409. doi: 10.1073/pnas.0408030101
Li, Z., Okamoto, K., Hayashi, Y., and Sheng, M. (2004). The importance of Palmer, A. E., and Tsien, R. Y. (2006). Measuring calcium signaling using
dendritic mitochondria in the morphogenesis and plasticity of spines and genetically targetable fluorescent indicators. Nat. Protoc. 1, 1057–1065. doi: 10.
synapses. Cell 119, 873–887. doi: 10.1016/j.cell.2004.11.003 1038/nprot.2006.172
Liang, Y., Yuan, L. L., Johnston, D., and Gray, R. (2002). Calcium signaling at single Patron, M., Checchetto, V., Raffaello, A., Teardo, E., Vecellio Reane, D.,
mossy fiber presynaptic terminals in the rat hippocampus. J. Neurophysiol. 87, Mantoan, M., et al. (2014). MICU1 and MICU2 finely tune the mitochondrial
1132–1137. Ca2+ uniporter by exerting opposite effects on MCU activity. Mol. Cell 53,
Llano, I., Gonzalez, J., Caputo, C., Lai, F. A., Blayney, L. M., Tan, Y. P., et al. 726–737. doi: 10.1016/j.molcel.2014.01.013
(2000). Presynaptic calcium stores underlie large-amplitude miniature IPSCs Perocchi, F., Gohil, V. M., Girgis, H. S., Bao, X. R., McCombs, J. E., Palmer, A. E.,
and spontaneous calcium transients. Nat. Neurosci. 3, 1256–1265. doi: 10. et al. (2010). MICU1 encodes a mitochondrial EF hand protein required for
1038/81781 Ca2+ uptake. Nature 467, 291–296. doi: 10.1038/nature09358
Lüscher, C., and Malenka, R. C. (2012). NMDA receptor-dependent long-term Plovanich, M., Bogorad, R. L., Sancak, Y., Kamer, K. J., Strittmatter, L., Li, A. A.,
potentiation and long-term depression (LTP/LTD). Cold Spring Harb Perspect. et al. (2013). MICU2, a paralog of MICU1, resides within the mitochondrial
Biol. 4:a005710. doi: 10.1101/cshperspect.a005710 uniporter complex to regulate calcium handling. PLoS One 8:e55785. doi: 10.
Ly, C. V., and Verstreken, P. (2006). Mitochondria at the synapse. Neuroscientist 1371/journal.pone.0055785
12, 291–299. doi: 10.1177/1073858406287661 Poburko, D., Santo-Domingo, J., and Demaurex, N. (2011). Dynamic regulation of
Malenka, R. C., and Nicoll, R. A. (1999). Long-term potentiation–a decade of the mitochondrial proton gradient during cytosolic calcium elevations. J. Biol.
progress? Science 285, 1870–1874. doi: 10.1126/science.285.5435.1870 Chem. 286, 11672–11684. doi: 10.1074/jbc.M110.159962
Mallilankaraman, K., Doonan, P., Cárdenas, C., Chandramoorthy, H. C., Raffaello, A., De Stefani, D., Sabbadin, D., Teardo, E., Merli, G., Picard, A., et al.
Müller, M., Miller, R., et al. (2012). MICU1 is an essential gatekeeper for MCU- (2013). The mitochondrial calcium uniporter is a multimer that can include
mediated mitochondrial Ca2+ uptake that regulates cell survival. Cell 151, a dominant-negative pore-forming subunit. EMBO J. 32, 2362–2376. doi: 10.
630–644. doi: 10.1016/j.cell.2012.10.011 1038/emboj.2013.157
Marland, J. R., Hasel, P., Bonnycastle, K., and Cousin, M. A. (2016). Mitochondrial Raymond, C. R., and Redman, S. J. (2002). Different calcium sources are narrowly
calcium uptake modulates synaptic vesicle endocytosis in central nerve tuned to the induction of different forms of LTP. J. Neurophysiol. 88, 249–255.
terminals. J. Biol. Chem. 291, 2080–2086. doi: 10.1074/jbc.M115.686956 Reyes, M., and Stanton, P. K. (1996). Induction of hippocampal long-term
Mathew, S. S., and Hablitz, J. J. (2008). Calcium release via activation of depression requires release of Ca2+ from separate presynaptic and postsynaptic
presynaptic IP3 receptors contributes to kainate-induced IPSC facilitation in intracellular stores. J. Neurosci. 16, 5951–5960.
rat neocortex. Neuropharmacology 55, 106–116. doi: 10.1016/j.neuropharm. Rizzuto, R., Brini, M., Murgia, M., and Pozzan, T. (1993). Microdomains with
2008.05.005 high Ca2+ close to IP3-sensitive channels that are sensed by neighboring
Mattson, M. P., Gleichmann, M., and Cheng, A. (2008). Mitochondria in mitochondria. Science 262, 744–747. doi: 10.1126/science.8235595
neuroplasticity and neurological disorders. Neuron 60, 748–766. doi: 10.1016/j. Rizzuto, R., De Stefani, D., Raffaello, A., and Mammucari, C. (2012). Mitochondria
neuron.2008.10.010 as sensors and regulators of calcium signalling. Nat. Rev. Mol. Cell Biol. 13,
Mellor, J., and Nicoll, R. A. (2001). Hippocampal mossy fiber LTP is independent 566–578. doi: 10.1038/nrm3412
of postsynaptic calcium. Nat. Neurosci. 4, 125–126. doi: 10.1038/83941 Rizzuto, R., and Pozzan, T. (2006). Microdomains of intracellular Ca2+ : molecular
Minta, A., Kao, J. P., and Tsien, R. Y. (1989). Fluorescent indicators for cytosolic determinants and functional consequences. Physiol. Rev. 86, 369–408. doi: 10.
calcium based on rhodamine and fluorescein chromophores. J. Biol. Chem. 264, 1152/physrev.00004.2005
8171–8178. Rizzuto, R., Simpson, A. W., Brini, M., and Pozzan, T. (1992). Rapid changes of
Mironov, S. L., and Symonchuk, N. (2006). ER vesicles and mitochondria move mitochondrial Ca2+ revealed by specifically targeted recombinant aequorin.
and communicate at synapses. J. Cell Sci. 119, 4926–4934. doi: 10.1242/jcs. Nature 358, 325–327. doi: 10.1038/358325a0
03254 Robert, V., Gurlini, P., Tosello, V., Nagai, T., Miyawaki, A., Di Lisa, F., et al. (2001).
Miyata, M., Finch, E. A., Khiroug, L., Hashimoto, K., Hayasaka, S., Oda, S. I., Beat-to-beat oscillations of mitochondrial [Ca2+ ] in cardiac cells. EMBO J. 20,
et al. (2000). Local calcium release in dendritic spines required for long- 4998–5007. doi: 10.1093/emboj/20.17.4998
term synaptic depression. Neuron 28, 233–244. doi: 10.1016/s0896-6273(00) Rodriguez-Garcia, A., Rojo-Ruiz, J., Navas-Navarro, P., Aulestia, F. J., Gallego-
00099-4 Sandin, S., Garcia-Sancho, J., et al. (2014). GAP, an aequorin-based fluorescent
Nagai, T., Sawano, A., Park, E. S., and Miyawaki, A. (2001). Circularly permuted indicator for imaging Ca2+ in organelles. Proc. Natl. Acad. Sci. U S A 111,
green fluorescent proteins engineered to sense Ca2+ . Proc. Natl. Acad. Sci. 2584–2589. doi: 10.1073/pnas.1316539111
U S A 98, 3197–3202. doi: 10.1073/pnas.051636098 Rose, T., Goltstein, P. M., Portugues, R., and Griesbeck, O. (2014). Putting a
Nagase, T., Ito, K. I., Kato, K., Kaneko, K., Kohda, K., Matsumoto, M., et al. finishing touch on GECIs. Front. Mol. Neurosci. 7:88. doi: 10.3389/fnmol.2014.
(2003). Long-term potentiation and long-term depression in hippocampal CA1 00088
neurons of mice lacking the IP(3) type 1 receptor. Neuroscience 117, 821–830. Rowland, K. C., Irby, N. K., and Spirou, G. A. (2000). Specialized synapse-
doi: 10.1016/s0306-4522(02)00803-5 associated structures within the calyx of Held. J. Neurosci. 20, 9135–9144.
Sancak, Y., Markhard, A. L., Kitami, T., Kovács-Bogdán, E., Kamer, K. J., Tang, Y., and Zucker, R. S. (1997). Mitochondrial involvement in post-tetanic
Udeshi, N. D., et al. (2013). EMRE is an essential component of the potentiation of synaptic transmission. Neuron 18, 483–491. doi: 10.1016/s0896-
mitochondrial calcium uniporter complex. Science 342, 1379–1382. doi: 10. 6273(00)81248-9
1126/science.1242993 Thayer, S. A., and Miller, R. J. (1990). Regulation of the intracellular free calcium
Sandler, V. M., and Barbara, J. G. (1999). Calcium-induced calcium release concentration in single rat dorsal root ganglion neurones in vitro. J. Physiol.
contributes to action potential-evoked calcium transients in hippocampal CA1 425, 85–115. doi: 10.1113/jphysiol.1990.sp018094
pyramidal neurons. J. Neurosci. 19, 4325–4336. Tian, L., Hires, S. A., and Looger, L. L. (2012). Imaging neuronal activity
Schneggenburger, R., and Neher, E. (2005). Presynaptic calcium and control of with genetically encoded calcium indicators. Cold Spring Harb. Protoc. 2012,
vesicle fusion. Curr. Opin. Neurobiol. 15, 266–274. doi: 10.1016/j.conb.2005. 647–656. doi: 10.1101/pdb.top069609
05.006 Unni, V. K., Zakharenko, S. S., Zablow, L., DeCostanzo, A. J., and Siegelbaum, S. A.
Schon, E. A., and Przedborski, S. (2011). Mitochondria: the next (2004). Calcium release from presynaptic ryanodine-sensitive stores is required
(neurode)generation. Neuron 70, 1033–1053. doi: 10.1016/j.neuron.2011. for long-term depression at hippocampal CA3-CA3 pyramidal neuron
06.003 synapses. J. Neurosci. 24, 9612–9622. doi: 10.1523/JNEUROSCI.5583-03.2004
Scullin, C. S., and Partridge, L. D. (2010). Contributions of SERCA pump and Verkhratsky, A. (2005). Physiology and pathophysiology of the calcium store
ryanodine-sensitive stores to presynaptic residual Ca2+ . Cell Calcium 47, in the endoplasmic reticulum of neurons. Physiol. Rev. 85, 201–279. doi: 10.
326–338. doi: 10.1016/j.ceca.2010.01.004 1152/physrev.00004.2004
Scullin, C. S., Wilson, M. C., and Partridge, L. D. (2010). Developmental changes in Verstreken, P., Ly, C. V., Venken, K. J., Koh, T. W., Zhou, Y., and Bellen, H. J.
presynaptic Ca2+ clearance kinetics and synaptic plasticity in mouse Schaffer (2005). Synaptic mitochondria are critical for mobilization of reserve pool
collateral terminals. Eur. J. Neurosci. 31, 817–826. doi: 10.1111/j.1460-9568. vesicles at Drosophila neuromuscular junctions. Neuron 47, 365–378. doi: 10.
2010.07137.x 1016/j.neuron.2005.06.018
Segal, M., and Korkotian, E. (2015). Roles of calcium stores and store- Voelker, D. R. (1990). Characterization of phosphatidylserine synthesis and
operated channels in plasticity of dendritic spines. Neuroscientist doi: 10. translocation in permeabilized animal cells. J. Biol. Chem. 265, 14340–14346.
1177/1073858415613277 [Epub ahead of print]. Vos, M., Lauwers, E., and Verstreken, P. (2010). Synaptic mitochondria in synaptic
Sharma, G., and Vijayaraghavan, S. (2003). Modulation of presynaptic store transmission and organization of vesicle pools in health and disease. Front.
calcium induces release of glutamate and postsynaptic firing. Neuron 38, Synaptic Neurosci. 2:139. doi: 10.3389/fnsyn.2010.00139
929–939. doi: 10.1016/s0896-6273(03)00322-2 Wang, S. S., Denk, W., and Häusser, M. (2000). Coincidence detection in single
Sharp, A. H., McPherson, P. S., Dawson, T. M., Aoki, C., Campbell, K. P., dendritic spines mediated by calcium release. Nat. Neurosci. 3, 1266–1273.
and Snyder, S. H. (1993). Differential immunohistochemical localization of doi: 10.1038/81792
inositol 1,4,5-trisphosphate- and ryanodine-sensitive Ca2+ release channels in Wang, X., and Schwarz, T. L. (2009). The mechanism of Ca2+ -dependent
rat brain. J. Neurosci. 13, 3051–3063. regulation of kinesin-mediated mitochondrial motility. Cell 136, 163–174.
Sharp, A. H., Nucifora, F. C. Jr., Blondel, O., Sheppard, C. A., Zhang, C., doi: 10.1016/j.cell.2008.11.046
Snyder, S. H., et al. (1999). Differential cellular expression of Wang, G. J., and Thayer, S. A. (2002). NMDA-induced calcium loads recycle
isoforms of inositol 1,4,5-triphosphate receptors in neurons and glia across the mitochondrial inner membrane of hippocampal neurons in culture.
in brain. J. Comp. Neurol. 406, 207–220. doi: 10.1002/(SICI)1096- J. Neurophysiol. 87, 740–749.
9861(19990405)406:2<207::AID-CNE6>3.0.CO;2-7 White, R. J., and Reynolds, I. J. (1995). Mitochondria and Na+ /Ca2+
Sheng, M., and Hoogenraad, C. C. (2007). The postsynaptic architecture of exchange buffer glutamate-induced calcium loads in cultured cortical neurons.
excitatory synapses: a more quantitative view. Annu. Rev. Biochem. 76, J. Neurosci. 15, 1318–1328.
823–847. doi: 10.1146/annurev.biochem.76.060805.160029 Wu, J., Prole, D. L., Shen, Y., Lin, Z., Gnanasekaran, A., Liu, Y., et al. (2014). Red
Shepherd, G. M., and Harris, K. M. (1998). Three-dimensional structure and fluorescent genetically encoded Ca2+ indicators for use in mitochondria and
composition of CA3→CA1 axons in rat hippocampal slices: implications endoplasmic reticulum. Biochem. J. 464, 13–22. doi: 10.1042/bj20140931
for presynaptic connectivity and compartmentalization. J. Neurosci. 18, Yang, S. N., Tang, Y. G., and Zucker, R. S. (1999). Selective induction of LTP and
8300–8310. LTD by postsynaptic [Ca2+ ]i elevation. J. Neurophysiol. 81, 781–787.
Shoshan-Barmatz, V., Zalk, R., Gincel, D., and Vardi, N. (2004). Subcellular Yeckel, M. F., Kapur, A., and Johnston, D. (1999). Multiple forms of LTP in
localization of VDAC in mitochondria and ER in the cerebellum. Biochim. hippocampal CA3 neurons use a common postsynaptic mechanism. Nat.
Biophys. Acta 1657, 105–114. doi: 10.1016/j.bbabio.2004.02.009 Neurosci. 2, 625–633. doi: 10.1038/10180
Slater, E. C., and Cleland, K. W. (1953). The effect of calcium on the respiratory Young, K. W., Bampton, E. T., Pinòn, L., Bano, D., and Nicotera, P. (2008).
and phosphorylative activities of heart-muscle sarcosomes. Biochem. J. 55, Mitochondrial Ca2+ signalling in hippocampal neurons. Cell Calcium 43,
566–590. doi: 10.1042/bj0550566 296–306. doi: 10.1016/j.ceca.2007.06.007
Solovyova, N., Veselovsky, N., Toescu, E. C., and Verkhratsky, A. (2002). Ca2+ Zhong, N., Beaumont, V., and Zucker, R. S. (2001). Roles for mitochondrial and
dynamics in the lumen of the endoplasmic reticulum in sensory neurons: direct reverse mode Na+ /Ca2+ exchange and the plasmalemma Ca2+ ATPase in
visualization of Ca2+ -induced Ca2+ release triggered by physiological Ca2+ post-tetanic potentiation at crayfish neuromuscular junctions. J. Neurosci. 21,
entry. EMBO J. 21, 622–630. doi: 10.1093/emboj/21.4.622 9598–9607.
Spacek, J., and Harris, K. M. (1997). Three-dimensional organization of smooth Zhang, H., Sun, S., Herreman, A., De Strooper, B., and Bezprozvanny, I.
endoplasmic reticulum in hippocampal CA1 dendrites and dendritic spines of (2010). Role of presenilins in neuronal calcium homeostasis. J. Neurosci. 30,
the immature and mature rat. J. Neurosci. 17, 190–203. doi: 10.1002/hipo.20238 8566–8580. doi: 10.1523/JNEUROSCI.1554-10.2010
Südhof, T. C. (2012). The presynaptic active zone. Neuron 75, 11–25. doi: 10. Zucker, R. S. (1989). Short-term synaptic plasticity. Annu. Rev. Neurosci. 12,
1016/j.neuron.2012.06.012 13–31. doi: 10.1146/annurev.neuro.12.1.13
Suzuki, J., Kanemaru, K., Ishii, K., Ohkura, M., Okubo, Y., and Iino, M. (2014).
Imaging intraorganellar Ca2+ at subcellular resolution using CEPIA. Nat. Conflict of Interest Statement: The authors declare that the research was
Commun. 5:4153. doi: 10.1038/ncomms5153 conducted in the absence of any commercial or financial relationships that could
Takechi, H., Eilers, J., and Konnerth, A. (1998). A new class of synaptic response be construed as a potential conflict of interest.
involving calcium release in dendritic spines. Nature 396, 757–760. doi: 10.
1038/25547 Copyright © 2016 Kwon, Hirabayashi and Polleux. This is an open-access article
Takei, K., Stukenbrok, H., Metcalf, A., Mignery, G. A., Südhof, T. C., distributed under the terms of the Creative Commons Attribution License (CC BY).
Volpe, P., et al. (1992). Ca2+ stores in Purkinje neurons: endoplasmic The use, distribution and reproduction in other forums is permitted, provided the
reticulum subcompartments demonstrated by the heterogeneous distribution original author(s) or licensor are credited and that the original publication in this
of the InsP3 receptor, Ca2+ -ATPase and calsequestrin. J. Neurosci. 12, journal is cited, in accordance with accepted academic practice. No use, distribution
489–505. or reproduction is permitted which does not comply with these terms.
Reviewed by:
Fabrice Ango,
University of Montpellier, France INTRODUCTION
Antonio Malgaroli,
Università Vita-Salute San Raffaele, One of the fundamental goals of neuroscience is to understand the generation of functional nervous
Italy system that underlies neural basis of behavior and cognition. Extensive research has attempted
*Correspondence: to interrogate the molecular and cellular mechanisms of synapse formation and functional
Mikyoung Park neural circuit development. Ever since it was proposed by Sydney Brenner in the mid 1960’s
[email protected]; (Brenner, 1974), the nematodes Caenorhabditis elegans (C. elegans) has been considered as an
[email protected] ideal model organism to study synaptic development and neural circuitry. The organism has
relatively simple nervous system, having 302 neurons and its neurochemistry and genetics are
Received: 25 April 2016 similar to those of mammals. Moreover, the complete structure and connectivity of C. elegans
Accepted: 17 June 2016 nervous system have been deciphered through genetic screens and reconstruction of electron
Published: 29 June 2016
micrographs (EM) of serial sections, which led to discovery of novel molecules important for
Citation: development and maintenance of functional synaptic connectivity (White et al., 1986). C. elegans
Hong J-H and Park M (2016)
with its transparent body was the first animal in which the green fluorescent protein (GFP)
Understanding Synaptogenesis and
Functional Connectome in C. elegans
was expressed (Chalfie et al., 1994). Combined with its stable expression of fluorescently tagged
by Imaging Technology. proteins (Mello et al., 1991; Frokjaer-Jensen et al., 2008), studies with C. elegans have made major
Front. Synaptic Neurosci. 8:18. contributions to our knowledge on neural development, axonal migration, and synapse formation.
doi: 10.3389/fnsyn.2016.00018 Recently, selective plane illumination microscopy (SPIM) techniques such as tiling light-sheet
SPIM (TLS-SPIM) (Fu et al., 2016) and inverted SPIM (iSPIM) Many studies using C. elegans have investigated the role of
(Wu et al., 2011) have been developed and utilized to achieve various proteins localized at active zone in synapse formation
high spatiotemporal resolution 3-dimensional live imaging of (Yeh et al., 2005; Watanabe et al., 2011). Classical EM
C. elegans embyos with no detectable phototoxicity, which could analysis has provided initial assessment of C. elegans synaptic
enable studies on synaptogenesis and axon guidance during components but its requirement for ultrathin sectioning of
embryogenesis in C. elegans. Another recent work adopting samples approximately 50 nm thickness (White et al., 1986)
complementation-activated light microscopy (CALM) in which limits the resolution and impairs detailed visualization of fine
proteins are conjugated with non-fluorescent split-fluorescent structures. The multifunctional synaptic scaffolding protein
proteins, which become to be fluorescent when complemented SYD-2/liprin-α is one of the key proteins identified to regulate
with synthetic peptides enabled single-molecule imaging with a synaptic development in C. elegans and Drosophila (Zhen
precision of 30 nm within synapses in live worms (Zhan et al., and Jin, 1999). The loss-of-function analysis on SYD-2/liprin-
2014). α and uncoordinated-10 (UNC-10)/Rab3-interacting molecule
Rapid developments of advanced imaging technologies have (RIM), which is another dense-projection components (Weimer
expanded our understanding of the molecular and cellular basis et al., 2006) revealed reduced vesicle recruitment at active zone
of synaptogenesis with great depth, taking a huge step closer (Stigloher et al., 2011; Kittelmann et al., 2013), and smaller dense-
to revealing functional neural connectome. Here, we discuss projection due to loss of SYD-2/liprin-α function (Kittelmann
on the synaptogenesis in presynaptic active zones revealed by et al., 2013) unlike the finding showing an expanded dense-
both conventional and advanced imaging set-ups and review projection (Zhen and Jin, 1999). One suggested explanation
recent work utilizing advanced imaging technology to unravel for variability in syd-2 mutant synaptic ultrastructure is due to
the functional connectome of neural circuits. Rather than the differences in fixation procedure (Kittelmann et al., 2013).
dealing with the mechanistic aspects of synapse formation Nevertheless, it is certain that advanced and optimized imaging
and neural circuits development, this review will mainly technique led to identification of regulatory proteins to retain
focus on how synaptic ultrastructure, synaptic formation, and synaptic vesicle at active zone.
functional neural connectome have been sophisticated by the A method which comprises of correlative fluorescence
advanced imaging technology. For more in-depth reviews on the electron microscopy was developed and optimized to observe
mechanism of synaptogenesis, synaptic specificity, and neural the nanoscopic localization of SYD-2/liprin-α in C. elegans
circuits development, see Campbell et al. (2015), Cherra and Jin active zone (Watanabe et al., 2011). The technique employed
(2015), Jin (2015), Zhen and Samuel (2015), Yogev and Shen both stimulated emission depletion (STED) microscopy and
(2014), Chia et al. (2013), and Park and Shen (2012). photoactivated localization microscopy (PALM) on ultrathin
sections for protein localization at super-resolution nanoscale
level and subsequently correlate the protein localization with
IMAGING SYNAPSE ASSEMBLY ultrastructures by electron microscope. The localization of
SYD-2/liprin-α to the C. elegans presynaptic dense-projection
Chemical synapses are specialized intercellular junctions with observed by this technique (Watanabe et al., 2011) was consistent
two apposed compartments, the pre-synaptic terminal and with the earlier finding from the immunoelectron micrograph
the postsynaptic target, and the synaptic cleft which is about (Yeh et al., 2005) but the result was more advanced to provide
20 nm gap between the pre- and postsynapses (Cowan and the precise localization of the proteins in small and dense
Kandel, 2001). Proper organization of pre- and postsynaptic structures likely within the synapse at the level of nanoscale
components with precise regulation underlies formation of super-resolution.
functional synapses. For the past decades, tremendous details In addition, studies using advanced EM tomography of
regarding the morphology and assembly of C. elegans synaptic 250 nm thick sections combined with high-pressure freezing
structure have been revealed with development of genetic tools (HPF) and freeze substitution (Stigloher et al., 2011; Kittelmann
and imaging technology. This section focuses on presynaptic et al., 2013) have resolved the highly complex structure of dense-
assembly and synaptic specificity revealed by genetically encoded projections at cholinergic neuromuscular junctions (NMJs) of
molecular tools and imaging technologies. C. elegans, revealing composition of building units forming bay-
like structures in which synaptic vesicles are docked to the
Presynaptic Active Zone Imaging active zone membrane. Furthermore, serial reconstruction of
The presynaptic compartment in C. elegans exhibits an overall HPF EM sections and EM tomography enabled the construction
structural organization similar to that in vertebrates, with of a high-resolution 3D model of presynaptic ultrastructure,
synaptic vesicles clustered in and around the electron-dense overcoming resolution limitation raised by the conventional EM
membrane structure called active zone known to serve as a and revealing a physical link between dense-projections and
major site of neurotransmitter release. Ultrastructural analysis synaptic vesicles within C. elegans presynaptic active zone.
have shown that, despite the variations among the appearances,
synapses of various organisms commonly display synaptic vesicle Presynaptic Assembly Imaging
docking and fusion at active zone that can be identified by darkly Cell type-specific tagging of synaptic proteins with fluorescent
stained membrane structures (Zhai and Bellen, 2004; Ackermann reporter has been a key reagent to study synaptogenesis and its
et al., 2015). regulation in C. elegans (Nonet, 1999; Shen and Bargmann, 2003;
Sieburth et al., 2005; Yeh et al., 2005). Hierarchical assembly to ectopic localizations of presynaptic components, including
of presynaptic active zone was observed in C. elegans HSNL RAB-3, SYD-1, SYD-2/liprin-α, GIT, and ELKS-1/ERC/CAST
synapses by fluorescently labeling the multiple active zone in the HSNL regions where the secondary vulval epithelial cells
proteins and expressing them in the various mutant animals made contacts (Figure 1A). This supported the idea that along
(Patel et al., 2006). Fluorescent protein fused with a synaptic with SYG-2/Nephrin as an upstream signal of SYG-1/Neph1,
vesicle-associated protein RAB-3 visualized synaptic vesicle SYG-1/Neph1 defines the presynaptic localization and is
clusters and confirmed the presynaptic localization of various sufficient to recruit presynaptic components, including the two
active zone components, including SYD-1, SYD-2/liprin-α, key scaffold molecules SYD-1 and SYD-2/liprin-α (Zhen and Jin,
ELKS-1/ERC/CAZ-associated structural protein (CAST), GIT, 1999) to the regions defined by its localization (Patel et al., 2006;
and SAD-1 kinase in the HSNL synapses (Patel et al., 2006). Figure 1B).
Altering the location of SYG-1/Neph1 by ectopically expressing In syd-1 and syd-2 mutants, the presynaptic components,
SYG-2/Nephrin in the secondary vulval epithelial cells, led including RAB-3, ELKS-1/ERC/CAST, GIT, SAD-1,
FIGURE 1 | Synaptic specificity regulated by non-neuronal factors. (A) Synaptic connectivity of neurons and muscles associated in the egg-laying circuit of C.
elegans. HSNL forms synapses with vulval muscle 2 (vm2) and ventral cord (VC) motor neurons, VC4 and VC5 specifically to the regions immediately adjacent to the
primary epithelial cells (1◦ ) which secretes SYG-2/Nephrin. Mutations in SYG-1/Neph1 or SYG-2/Nephrin disrupt synaptic specificity of HSNL and cause ectopic
synapse formation with select body wall muscle (BWM). Ectopic positioning of SYG-2/Nephrin to the secondary epithelial cells (2◦ ) recruited SYG-1/Neph1 to HSNL
near the secondary epithelial cells, which was shown to be sufficient to form synapses ectopically at the sites where SYG-1/Neph1 is recruited (yellow circles). (B)
Pathway for HSNL synapse assembly. SYG-2/Nephrin ensures proper localization of SYG-1/Neph1 which defines presynaptic localization of the active zone proteins.
ELKS-1/ERC/CAST could function redundantly with SYD-1 or other unidentified presynaptic proteins that positively regulate synapse assembly (gray lines). In the
presence of SYD-1, the SYD-2/liprin-α and ELKS-1/ERC/CAST interaction was enhanced (red arrow). In the presence of RSY-1, SYD-1, and ELKS-1/ERC/CAST
interaction is weakened (solid green), suggesting RSY-1 as a negative regulator in the HSNL presynaptic assembly process likely by weakening the SYD-2/liprin-α and
ELKS-1/ERC/CAST interaction (dotted green) indirectly through the RSY-1 and SYD-1 interaction. Plain lines indicate biochemical interactions. (C) Synaptic
connectivity of AIY and RIA interneurons regulated by ventral cephalic sheath cells (CEPshV) at C. elegans nerve ring. Synapses between AIY and RIA are formed en
passent as they are ensheathed in zone 2 by CEPshV, which secretes UNC-6/Netrin that regulates UNC-40/DCC activity in AIY. Abnormal distend positioning of
CEPshV toward zone 1 causes ectopic localizations of both presynapses (red circles) and UNC-40/DCC (purple triangles) in zone 1 of AIY. (D) Pathways for AIY and
RIA connectivity. CEPshV secretes UNC-6/Netrin, which regulates both positioning of presynapses in AIY and axon guidance of postsynaptic RIA through
UNC-40/DCC activity to the location specified by CEPshV.
UNC-57/endophilin, and SNN-1/synapsin-1 were failed to Inaki et al., 2007) or from a guidepost cell (Christopherson
be assembled, identifying those presynaptic components et al., 2005; Colon-Ramos et al., 2007). In C. elegans, synaptic
as downstream molecules of SYD-1 and SYD-2/liprin-α in the contacts are usually formed en passant, in which synapses are
active zone assembly process (Patel et al., 2006). Gain-of-function formed along the adjacent processes but not its terminus (White
mutation (Dai et al., 2006) or overexpression of SYD-2/liprin-α et al., 1986). Synaptic specificity studies in C. elegans have been
(Patel et al., 2006) in syd-1 mutants completely restored the accelerated upon the development of the expression tool of
synaptic accumulation of SNB-1/synaptobrevin, whereas the fluorescently tagged proteins in specific cell types driven by cell
SYD-1 overexpression in syd-2 mutants was not sufficient to type-specific promoters, which enabled researchers to specifically
induce the rescue effect (Patel et al., 2006), illustrating the SYD-1 label pre-, postsynapses, and neighboring guidepost cells (Nonet,
and SYD-2/liprin-α mediated presynaptic assembly with the 1999; Shen and Bargmann, 2003; Grunwald et al., 2004; Francis
SYD-1 as an upstream of SYD-2/liprin-α (Figure 1B). Although et al., 2005; Sieburth et al., 2005; Yeh et al., 2005; Hoerndli et al.,
the loss of ELKS-1 function by itself did not induce apparent 2013).
defects in synapse assembly in C. elegans HSNL synapses (Dai A specific synaptic connectivity between amphid interneuron
et al., 2006; Patel et al., 2006), synapse formation in the syd-2 Y (AIY) and ring interneuron A (RIA) in C. elegans nerve ring,
gain-of-function and syd-1 double mutants exhibited a high considered as brain of the animal, was fluorescently visualized
dependency on ELKS-1 expression (Dai et al., 2006), suggesting by expressing presynaptic RAB-3 in AIY and postsynaptic
that ELKS-1 functions redundantly with SYD-1 or other glutamate receptors GLR-1 in RIA (Colon-Ramos et al., 2007;
presynaptic proteins that positively regulate synapse assembly Shao et al., 2013) (Figure 1C). The localization of synaptic
(Figure 1B). connectivity between AIY and RIA has shown to be restricted
Regulator of synaptogenesis-1 (RSY-1) was cloned as a in the zone 2 of AIY axon (Figure 1C) and such specificity
negative regulator of synapse formation for its deletion mutants is achieved by activation of both UNC-6/Netrin, a well-known
to lead to extra synapse formation and exhibit increased axon guidance molecule that is exclusively expressed by glia-
accumulation of SNB-1/synaptobrevin at presynaptic sites in like ventral cephalic sheath cells (CEPshV) (Wadsworth et al.,
the HSNL (Patel and Shen, 2009). A single-cell in situ 1996) and the netrin receptor UNC-40/Deleted in Colorectal
protein-protein interaction assay revealed enhanced interaction Cancer (DCC) (Colon-Ramos et al., 2007), supporting the idea
between SYD-2/liprin-α and ELKS-1/ERC/CAST in presence that secreted molecules from glia govern synaptic specificity.
of SYD-1 while direct interaction between SYD-1 and ELKS- Confocal microscopy revealed the projection of the CEPshV
1/ERC/CAST is weakened in presence of RSY-1, suggesting processes with respect to the region of innervation between
RSY-1 as a negative regulator of C. elegans HSNL synapse AIY and RIA (Figure 1C). Loss-of-function in either UNC-
assembly likely by weakening the SYD-2/liprin-α and ELKS- 34/enabled, a regulator of the actin cytoskeleton (Colon-Ramos
1/ERC/CAST interaction indirectly through its interaction with et al., 2007) or circuit maintenance abnormal protein (CIMA-
SYD-1 (Figure 1B). Together, presynaptic differentiation at 1), a regulator of synaptic maintenance in C. elegans (Shao
C. elegans HSNL synapses was initiated by SYG-1/Neph1, et al., 2013), caused morphological alterations in CEPshV which
a synaptic specificity molecule that defines the location of migrated toward further posteriorly to ensheath AIY axon in
presynaptic sites along the HSNL axon, leading to activate zone 1 (Figure 1C). Morphological alterations in CEPshV led
the presynaptic assembly process by recruiting the two key to ectopic localization of both UNC-40/DCC and presynaptic
scaffolding proteins SYD-1 and SYD-2/liprin-α. SYD-2/liprin- components in zone 1 (Figure 1C) due to the existence of UNC-
α-centered assembly of presynaptic components was achieved 6/Netrin secreted from CEPshV in zone 1 (Colon-Ramos et al.,
through the inter-communications among positive (SYD- 2007). The process of RIA in unc-34 mutants also abnormally
1 and ELKS-1/ERC/CAST) and negative (RSY-1) regulators migrated toward zone 1 where the ectopic synapses were formed
(Figure 1B). (Figure 1C). Together, UNC-40/DCC plays two independent
roles in each neuron, which are positioning of presynapses in
AIY and axon guidance of postsynaptic RIA to the location
IMAGING SYNAPTIC SPECIFICITY specified by CEPshV (Figure 1D). These findings further support
the model of non-neuronal contribution to the regulation of
During the event of synapse formation, a precise apposition precise localization of synaptogenesis.
between the presynaptic release sites and postsynaptic receptors Earlier than the AIY-RIA synaptic specificity study, the
must be accomplished to ensure a rapid neurotransmitter C. elegans egg-laying circuit, which is predominantly innervated
release and reliable synaptic response. Neurons can select by the two hermaphrodite-specific motor neurons (HSNs), HSNL
subpopulations of neurons they form synapses onto and can also and HSNR, and the two ventral cord (VC) motor neurons, VC4
select the defined specific subcellular sites to establish synapses and VC5 has been reported to be regulated by non-neuronal
(Akins and Biederer, 2006; White, 2007; Margeta and Shen, factor. HSNL and HSNR synapse onto vulval muscle cells and
2010). Such synaptic specificity is achieved by trans-synaptic onto the VC4 and VC5 neurons, while VC4 and VC5 neurons also
adhesion between pre- and postsynaptic neurons (Yamagata synapse onto the vulval muscle cells. Despite the direct contact
et al., 2002; Graf et al., 2004; Choe et al., 2006), adhesion between HSN and VC processes, synapses formed between these
between the presynaptic neuron and a guidepost cell (Shen cells are only restricted to the regions adjacent to the vulva
and Bargmann, 2003; Shen et al., 2004), molecules secreted (White et al., 1986) (Figure 1A). The specific positioning of
from pre- or postsynaptic neurons (Umemori et al., 2004; synapses and the recognition between HSNL and its target
were determined by adjacent vulva epithelial guidepost cells that light-sensitive proteins such as channelrhodopsins (Nagel et al.,
express SYG-2/Nephrin. SYG-2/Nephrin interacts with SYG- 2003, 2005), halorhodopsins (Han and Boyden, 2007; Zhang
1/Neph1 expressed in the HSNL, to recruit SYG-1/Neph1 to the et al., 2007; Husson et al., 2012b), and archaerhodopsins (Ihara
site along the HSNL axon where presynaptic sites are developed et al., 1999) as optogenetic actuators to either activate or inhibit
(Shen and Bargmann, 2003; Shen et al., 2004) (Figure 1A). neuronal activity via light and genetically encoded sensors such
More recently, introduction of the GFP reconstitution as GCaMP calcium indicator (Tian et al., 2009) and Clomeleon
across synaptic partners (GRASP) developed in C. elegans chloride indicator (Kuner and Augustine, 2000; Berglund et al.,
has overcome the challenges addressed by labor-intensive 2006) as optogenetic sensors to monitor responses to the synaptic
conventional EM analysis and increased the spatial resolution inputs. This section will discuss various experimental imaging
to visualize the pre- and postsynaptic contacts. GRASP is based approaches to interrogate the neural connection using the C.
on functional complementation between two non-fluorescent elegans nervous system.
split-GFP fragments separately expressed in the pre- and The initial optogenetics was applied to manipulate the
postsynaptic neurons, which label synapses between two cells behavior of C. elegans (Nagel et al., 2005). Expression of
of close proximity in living animals (Feinberg et al., 2008). Channelrhodopsin-2 (ChR2), a blue light-gated depolarizing
Using GRASP, specific visualization of synaptic contacts between cation channel used to activate neural activity in C. elegans
AIY and RIA was observed with high spatial resolution (Shao body muscles caused blue light-evoked contractions (Nagel et al.,
et al., 2013). In addition, GRASP revealed restricted synaptic 2003, 2005), whereas expression of NpHR, a yellow light-gated
localization between AIY and CEPshV (Shao et al., 2013), which hyperpolarizing chloride ion pump applied to inhibit neural
is consistent with the published EM data (White et al., 1986). activity in C. elegans muscle cells caused an extension of the
Formation of ectopic synapses between AIY and CEPshV due worm’s body length and locomotion defects by whole-field
to morphological alteration in CEPshV was confirmed as well illumination of yellow-green light (Zhang et al., 2007; Husson
(Shao et al., 2013) (Figure 1C). GRASP application has also et al., 2012b). Aside from NpHR, a yellow-green light-sensitive
confirmed the SYG-1/Neph1 and SYG-2/Nephrin as synaptic archaerhodopsin-3 (Arch) (Ihara et al., 1999) and a green-
specificity regulators of HSN synapses with vulval muscles and blue light-sensitive Mac (Waschuk et al., 2005) have also been
VC neurons. Analyzing GRASP fluorescence in wild-type and expressed in C. elegans and induced a stronger optical silencing
syg-1 or syg-2 mutants recapitulated the synaptic connectivity effect than NpHR likely due to efficient protein trafficking to
of HSN neurons (Feinberg et al., 2008) (Figure 1A). Besides the plasma membrane (Chow et al., 2010; Husson et al., 2012b).
the C. elegans nervous system, the GRASP has also been widely The simultaneous use of Arch and Mac enabled inhibition
adapted by other model systems, such as Drosophila (Gordon and of two different neuronal subpopulations, depending on the
Scott, 2009; Gong et al., 2010) mouse (Kim et al., 2012; Yamagata illuminating lights used.
and Sanes, 2012) and the cultured hippocampal neuronal system Light-sensitive probes expressed in C. elegans in vivo are
(Tsetsenis et al., 2014). Lately, newly modified GRASP strategies, mostly under the control of a promoter sequence. However,
involving activity-dependent synaptic GRASP and multi-color promoter-driven single cell expression of optogenetic protein
fluorescence reconstitution across synapses (X-RASP) have been is challenging to achieve due to the lack of single cell-specific
validated in Drosophila, allowing preferential labeling of active promoter and instead proteins are diversely expressed, eliciting
synapses and multi-color labeling of active synapses in one robust behavioral responses upon whole-field illumination
animal (Macpherson et al., 2015; Li et al., 2016). Continuous (Husson et al., 2013). Although it may be useful for inspecting
development of GRASP shows the potential to expand the utility a novel optogenetic protein, optical manipulation of individual
of GRASP to identify and map synaptic connectivity of neural neurons needs to be accomplished in order to obtain insights
circuits in the living animal with high resolution. into individual contribution by single neurons in functional
connectivity. To this aim, new methods have been adapted
in C. elegans to drive selective optical manipulation, either
IMAGING FUNCTIONAL NEURAL by genetically modulated single cell-specific expression of
CIRCUITS optogenetic protein (Ezcurra et al., 2011; Schmitt et al., 2012; Cho
and Sternberg, 2014; Guo et al., 2015) or selective illumination of
An underlying goal of neuroscience is to understand the neural target neurons with a high spatial and temporal resolution (Guo
connectome that are responsible for synaptic function and et al., 2009; Leifer et al., 2011; Stirman et al., 2011; Husson et al.,
neuronal basis of behavior. Anatomical structural connectome 2012a,b; Kocabas et al., 2012; Cohen et al., 2014; Luo et al., 2014;
of the whole nervous system of C. elegans, which has been fully Shipley et al., 2014; Trojanowski et al., 2014) in order to dissect
mapped by EM of serial sections (White et al., 1986), has served functional connections within the neural circuits (Table 1).
as a useful resource for researchers to study circuit function, Mainly adapted approach to specifically deliver light-sensitive
thus making the C. elegans nervous system as an excellent opsins to individual neurons of C. elegans restricts the opsin
model to investigate functional connectome of neural circuits. expression by genetic application using Cre or FLP recombinases
For the past decade, optogenetics has been widely adapted (Ezcurra et al., 2011; Schmitt et al., 2012; Cho and Sternberg,
to manipulate neural circuits and examine the corresponding 2014; Guo et al., 2015) (Figure 2). The recombinase-dependent
changes in synaptic function and behavior (Fang-Yen et al., gene expression is driven by a set of two promoters, a first
2012; Husson et al., 2013). Optogenetics uses genetically encoded promoter driving the expression of opsin conjugated with a
*Optogenetic manipulation driven by neuronal type-specific promoters rather than Cre/FLP recombinase application.
fluorophore along with or without a bicistronic fluorescent ASH to the interneurons AVA and AVD and the connections
reporter and a second promoter driving the expression of Cre between the interneurons RIM and AVA have been monitored
or FLP recombinase. In the first promoter-containing construct, (Guo et al., 2009; Table 1).
a transcription termination sequence flanked by recombinase Improvement in microscopic analysis and optogenetic
recognition sequences, loxP or FRT that are recognized by Cre illumination system allowed manipulation of neural activity in a
or FLP recombinase is enclosed in front of opsin. The Cre or freely behaving C. elegans with a high spatiotemporal resolution,
FLP recombinase-mediated recombination of loxP or FRT sites providing an in-depth analysis on functional neural circuits
excised the stop sequence and allows conditional expression underlying behavior at a single-cell level. A modified three-
of opsin only in the target cell where both promoters are panel liquid crystal display (3-LCD) projector for simultaneous
active (Husson et al., 2013) (Figure 2). Using Cre and FLP multicolor illumination and a motorized X-Y stage for keeping
system, ChR2 were specifically expressed in PVC interneurons the unrestrained worm centered in the camera’s field of view
which evoked forward locomotion and in AVA interneuron and with a standard inverted epifluorescence microscope were
ASH sensory neurons which evoked backward-movement upon systemized (Stirman et al., 2011; Husson et al., 2012b) and
photostimulation (Ezcurra et al., 2011; Schmitt et al., 2012) the Colbert system was equipped to control locomotion and
(Table 1). Further effort to isolate exclusive expression of the behavior in real time (Leifer et al., 2011; Luo et al., 2014;
light-sensitive proteins in a single cell (Ezcurra et al., 2011) would Shipley et al., 2014). Spatial regulation of optical illumination
need to define the role of individual single neurons in functional is controlled either by estimating the coordinates of targeted
neural circuits. cells using the machine-vision algorithms (Leifer et al., 2011;
Instead of using genetically generated system and whole-field Trojanowski et al., 2014) or by calculating the anterior-posterior
illumination, spatiotemporally patterned illumination of neurons (A-P) axis (Stirman et al., 2011). Both systems have been
expressing light-sensitive optogenetic proteins in immobilized instrumental in defining neural coding of several behaviors
C. elegans was used by for the first time in vivo using a digital in C. elegans linked to the motor circuit, avoidance circuit,
micromirror device (DMD) whose individual mirrors can be nociceptive circuit, chemotaxis circuit, and feeding circuits of
controlled independently to precisely determine the location freely moving worms (Table 1). Using AIY expressing ChR2 and
and size of the regions to be illuminated while simultaneously targeted illumination by the DMD technology, it was shown that
recording the calcium levels using a genetically encoded calcium optogenetic manipulation of AIY activity alone was sufficient
sensor, GCaMP to analyze the functional connections among to evoke chemotactic behavior in freely moving C. elegans, and
neurons. Combining the optogenetic actuator ChR2 and the was suggested that AIY is plausible to act as a control node
sensor GCaMP with the patterned illumination via a DMD for coordinating other taxis behaviors as well (Kocabas et al.,
technology, the functional connections from the sensory neuron 2012). Another report using the Colbert system equipped with
FIGURE 2 | Restricted expression of light-sensitive opsin mediated by Cre or FLP recombinases. Promoter 1-containing construct is designed to drive
expression of opsin with a fluorescent reporter. Promoter 2 drives expression of Cre or FLP recombinase. Conditional expression of opsin is mediated by the Cre or
FLP recombinases by removing a transcription termination sequence flanked by loxP or FRT only in target cell where the both promoters are active.
the DMD investigated an experience-dependent salt chemotaxis As genetically encoded fluorescent proteins have been rapidly
circuit. Optogenetic manipulation of neuronal activity of the developed for the past decades since the GFP was introduced
ASER sensory neuron expressing ChR2 was shown to be in the field, it is also expected that the number of optogenetic
connected to positive and negative chemotaxis in response to salt tools will rapidly increase to likely provide optogenetic proteins
concentrations, indicating that ASER sensory neuron encodes with different spectral properties (Zhang et al., 2008; Gradinaru
the perception of salt concentration and the memory of the et al., 2010) and ionic specificities (Han and Boyden, 2007;
chemotactic set point in a chemotaxis circuit of C. elegans (Luo Zhang et al., 2007) and help expand the understanding of
et al., 2014). In addition, optogenetic manipulations of specific synaptic function and neural circuits. During such processes,
pharyngeal neurons MC, M2, M4, and I1 in freely behaving it is confidently predicted that C. elegans will provide a
worms by adopting ChR2 for optical stimulation and Mac for systematic in vivo platform to test the optogenetic tools newly
optical silencing along with the DMD for targeted illumination developed and to ultimately apply to the synaptic function and
revealed a pharyngeal pumping/feeding circuit and identified the functional connectome studies. Together with the improvement
regulation of feeding rate by nicotinic and muscarinic receptors of fluorescent and optogenetic tools, continuous development in
through the pharyngeal neuronal network (Trojanowski et al., C. elegans imaging technology will promise a breakthrough in
2014). Furthermore, multispectral illumination (Stirman et al., deciphering functional neural connectome.
2011) enables simultaneous application of optical stimulation In addition to the monitoring and controlling of existing
and inhibition to an individual animal. Emerging studies have neuronal circuits via optogenetic applications and advanced
successfully facilitated multimodal optogenetic manipulation on microscopy systems as described in this review, it is very
C. elegans to independently excite different neurons in a single plausible to develop the ways to actively manipulate neural
worm (Erbguth et al., 2012; Husson et al., 2012b; Schild and circuits for instance by inserting new connections or removing
Glauser, 2015). existing connections, resulting in the reprogramming of neural
circuits. Indeed, a recent study on artificial modifications
PERSPECTIVES of neural circuits was reported in C. elegans by expressing
transgenically targeted heterologous connexin to insert
C. elegans is currently the best organism to study synapses and a new electrical synapse between normally unconnected
neuronal circuits because the connectivity of its 302 neurons neurons in intact animals, which resulted in altered salt
has been well-defined by serial reconstruction of EM (White taste and olfactory chemotaxis behavior (Rabinowitch
et al., 1986), the body is transparent, and it is a genetically et al., 2014). Conversely, laser ablation method can be
tractable animal model. C. elegans was one of the first organisms used to remove existing connection (Sulston and White,
that GFP was expressed to label protein (Chalfie et al., 1994), 1980; Bargmann et al., 1993; McIntire et al., 1993; Fang-
GRASP was utilized to visualize specific synaptic contacts Yen et al., 2012; Rabinowitch et al., 2013). Such artificial
(Feinberg et al., 2008), optogenetics was applied to manipulate modification of neural circuits not only help understand
behavior of live animals (Nagel et al., 2005), and more recently, fundamental functions of neuronal connectivity underlying
sonogenetics using low-pressure ultrasound was challenged to complex behavior but could also be applied to disease
activate specific ultrasonically sensitized neurons and modify brain circuits with the purpose of therapeutics at the circuit
locomotory behavior (Ibsen et al., 2015). level.
REFERENCES Fang-Yen, C., Gabel, C. V., Samuel, A. D., Bargmann, C. I., and Avery, L. (2012).
Laser microsurgery in Caenorhabditis elegans. Methods Cell Biol. 107, 177–206.
Ackermann, F., Waites, C. L., and Garner, C. C. (2015). Presynaptic active doi: 10.1016/B978-0-12-394620-1.00006-0
zones in invertebrates and vertebrates. EMBO Rep. 16, 923–938. doi: Feinberg, E. H., Vanhoven, M. K., Bendesky, A., Wang, G., Fetter, R. D., Shen, K.,
10.15252/embr.201540434 et al. (2008). GFP Reconstitution Across Synaptic Partners (GRASP) defines
Akins, M. R., and Biederer, T. (2006). Cell-cell interactions in synaptogenesis. Curr. cell contacts and synapses in living nervous systems. Neuron 57, 353–363. doi:
Opin. Neurobiol. 16, 83–89. doi: 10.1016/j.conb.2006.01.009 10.1016/j.neuron.2007.11.030
Bargmann, C. I., Hartwieg, E., and Horvitz, H. R. (1993). Odorant-selective Francis, M. M., Evans, S. P., Jensen, M., Madsen, D. M., Mancuso, J., Norman, K.
genes and neurons mediate olfaction in C. elegans. Cell 74, 515–527. doi: R., et al. (2005). The Ror receptor tyrosine kinase CAM-1 is required for ACR-
10.1016/0092-8674(93)80053-H 16-mediated synaptic transmission at the C. elegans neuromuscular junction.
Berglund, K., Schleich, W., Krieger, P., Loo, L. S., Wang, D., Cant, N. B., et al. Neuron 46, 581–594. doi: 10.1016/j.neuron.2005.04.010
(2006). Imaging synaptic inhibition in transgenic mice expressing the chloride Frokjaer-Jensen, C., Davis, M. W., Hopkins, C. E., Newman, B. J., Thummel, J. M.,
indicator, Clomeleon. Brain Cell Biol. 35, 207–228. doi: 10.1007/s11068-008- Olesen, S. P., et al. (2008). Single-copy insertion of transgenes in Caenorhabditis
9019-6 elegans. Nat. Genet. 40, 1375–1383. doi: 10.1038/ng.248
Brenner, S. (1974). The genetics of Caenorhabditis elegans. Genetics 77, 71–94. Fu, Q., Martin, B. L., Matus, D. Q., and Gao, L. (2016). Imaging multicellular
Campbell, J. C., Chin-Sang, I. D., and Bendena, W. G. (2015). Mechanosensation specimens with real-time optimized tiling light-sheet selective plane
circuitry in Caenorhabditis elegans: a focus on gentle touch. Peptides 68, illumination microscopy. Nat. Commun. 7:11088. doi: 10.1038/ncomms
164–174. doi: 10.1016/j.peptides.2014.12.004 11088
Chalfie, M., Tu, Y., Euskirchen, G., Ward, W. W., and Prasher, D. C. (1994). Green Gong, Z., Liu, J., Guo, C., Zhou, Y., Teng, Y., and Liu, L. (2010). Two pairs of
fluorescent protein as a marker for gene expression. Science 263, 802–805. doi: neurons in the central brain control Drosophila innate light preference. Science
10.1126/science.8303295 330, 499–502. doi: 10.1126/science.1195993
Cherra, S. J. III, and Jin, Y. (2015). Advances in synapse formation: forging Gordon, M. D., and Scott, K. (2009). Motor control in a Drosophila taste circuit.
connections in the worm. Wiley Interdiscip. Rev. Dev. Biol. 4, 85–97. doi: Neuron 61, 373–384. doi: 10.1016/j.neuron.2008.12.033
10.1002/wdev.165 Gradinaru, V., Zhang, F., Ramakrishnan, C., Mattis, J., Prakash, R., Diester, I.,
Chia, P. H., Li, P., and Shen, K. (2013). Cell biology in neuroscience: cellular et al. (2010). Molecular and cellular approaches for diversifying and extending
and molecular mechanisms underlying presynapse formation. J. Cell Biol. 203, optogenetics. Cell 141, 154–165. doi: 10.1016/j.cell.2010.02.037
11–22. doi: 10.1083/jcb.201307020 Graf, E. R., Zhang, X., Jin, S. X., Linhoff, M. W., and Craig, A. M. (2004). Neurexins
Cho, J. Y., and Sternberg, P. W. (2014). Multilevel modulation of a sensory induce differentiation of GABA and glutamate postsynaptic specializations via
motor circuit during C. elegans sleep and arousal. Cell 156, 249–260. doi: neuroligins. Cell 119, 1013–1026. doi: 10.1016/j.cell.2004.11.035
10.1016/j.cell.2013.11.036 Grunwald, M. E., Mellem, J. E., Strutz, N., Maricq, A. V., and Kaplan, J. M.
Choe, K. M., Prakash, S., Bright, A., and Clandinin, T. R. (2006). Liprin-alpha is (2004). Clathrin-mediated endocytosis is required for compensatory regulation
required for photoreceptor target selection in Drosophila. Proc. Natl. Acad. Sci. of GLR-1 glutamate receptors after activity blockade. Proc. Natl. Acad. Sci.
U.S.A. 103, 11601–11606. doi: 10.1073/pnas.0601185103 U.S.A. 101, 3190–3195. doi: 10.1073/pnas.0306156101
Chow, B. Y., Han, X., Dobry, A. S., Qian, X., Chuong, A. S., Li, M., et al. Guo, M., Wu, T. H., Song, Y. X., Ge, M. H., Su, C. M., Niu, W. P., et al.
(2010). High-performance genetically targetable optical neural silencing (2015). Reciprocal inhibition between sensory ASH and ASI neurons modulates
by light-driven proton pumps. Nature 463, 98–102. doi: 10.1038/nature nociception and avoidance in Caenorhabditis elegans. Nat. Commun. 6, 5655.
08652 doi: 10.1038/ncomms6655
Christopherson, K. S., Ullian, E. M., Stokes, C. C., Mullowney, C. E., Hell, J. Guo, Z. V., Hart, A. C., and Ramanathan, S. (2009). Optical interrogation
W., Agah, A., et al. (2005). Thrombospondins are astrocyte-secreted proteins of neural circuits in Caenorhabditis elegans. Nat. Methods 6, 891–896. doi:
that promote CNS synaptogenesis. Cell 120, 421–433. doi: 10.1016/j.cell.2004. 10.1038/nmeth.1397
12.020 Han, X., and Boyden, E. S. (2007). Multiple-color optical activation, silencing,
Cohen, E., Chatzigeorgiou, M., Husson, S. J., Steuer-Costa, W., Gottschalk, A., and desynchronization of neural activity, with single-spike temporal resolution.
Schafer, W. R., et al. (2014). Caenorhabditis elegans nicotinic acetylcholine PLoS ONE 2:e299. doi: 10.1371/journal.pone.0000299
receptors are required for nociception. Mol. Cell. Neurosci. 59, 85–96. doi: Hoerndli, F. J., Maxfield, D. A., Brockie, P. J., Mellem, J. E., Jensen, E., Wang, R.,
10.1016/j.mcn.2014.02.001 et al. (2013). Kinesin-1 regulates synaptic strength by mediating the delivery,
Colon-Ramos, D. A., Margeta, M. A., and Shen, K. (2007). Glia promote local removal, and redistribution of AMPA receptors. Neuron 80, 1421–1437. doi:
synaptogenesis through UNC-6 (netrin) signaling in C. elegans. Science 318, 10.1016/j.neuron.2013.10.050
103–106. doi: 10.1126/science.1143762 Husson, S. J., Costa, W. S., Wabnig, S., Stirman, J. N., Watson, J. D., Spencer, W. C.,
Cowan, W. M., and Kandel, E. R. (2001). “A brief history of synapses and synaptic et al. (2012a). Optogenetic analysis of a nociceptor neuron and network reveals
transmission,” in Synapses,eds T. C. Sudhof, W. M. Cowan, and C. F. Stevens ion channels acting downstream of primary sensors. Curr. Biol. 22, 743–752.
(Baltimore: The Johns Hopkins University Press), 1–88. doi: 10.1016/j.cub.2012.02.066
Dai, Y., Taru, H., Deken, S. L., Grill, B., Ackley, B., Nonet, M. L., et al. (2006). SYD- Husson, S. J., Gottschalk, A., and Leifer, A. M. (2013). Optogenetic manipulation
2 Liprin-alpha organizes presynaptic active zone formation through ELKS. Nat. of neural activity in C. elegans: from synapse to circuits and behaviour. Biol. Cell
Neurosci. 9, 1479–1487. doi: 10.1038/nn1808 105, 235–250. doi: 10.1111/boc.201200069
Erbguth, K., Prigge, M., Schneider, F., Hegemann, P., and Gottschalk, A. Husson, S. J., Liewald, J. F., Schultheis, C., Stirman, J. N., Lu, H., and Gottschalk,
(2012). Bimodal activation of different neuron classes with the spectrally red- A. (2012b). Microbial light-activatable proton pumps as neuronal inhibitors to
shifted channelrhodopsin chimera C1V1 in Caenorhabditis elegans. PLoS ONE functionally dissect neuronal networks in C. elegans. PLoS One 7:e40937. doi:
7:e46827. doi: 10.1371/journal.pone.0046827 10.1371/journal.pone.0040937
Ezcurra, M., Tanizawa, Y., Swoboda, P., and Schafer, W. R. (2011). Food sensitizes Ibsen, S., Tong, A., Schutt, C., Esener, S., and Chalasani, S. H. (2015). Sonogenetics
C. elegans avoidance behaviours through acute dopamine signalling. EMBO J. is a non-invasive approach to activating neurons in Caenorhabditis elegans. Nat.
30, 1110–1122. doi: 10.1038/emboj.2011.22 Commun. 6, 8264. doi: 10.1038/ncomms9264
Ihara, K., Umemura, T., Katagiri, I., Kitajima-Ihara, T., Sugiyama, Y., Kimura, Y., Schild, L. C., and Glauser, D. A. (2015). Dual Color Neural Activation and Behavior
et al. (1999). Evolution of the archaeal rhodopsins: evolution rate changes by Control with Chrimson and CoChR in Caenorhabditis elegans. Genetics 200,
gene duplication and functional differentiation. J. Mol. Biol. 285, 163–174. doi: 1029–1034. doi: 10.1534/genetics.115.177956
10.1006/jmbi.1998.2286 Schmitt, C., Schultheis, C., Pokala, N., Husson, S. J., Liewald, J. F.,
Inaki, M., Yoshikawa, S., Thomas, J. B., Aburatani, H., and Nose, A. (2007). Wnt4 Bargmann, C. I., et al. (2012). Specific expression of channelrhodopsin-
is a local repulsive cue that determines synaptic target specificity. Curr. Biol. 17, 2 in single neurons of Caenorhabditis elegans. PLoS ONE 7:e43164. doi:
1574–1579. doi: 10.1016/j.cub.2007.08.013 10.1371/journal.pone.0043164
Jin, Y. (2015). Unraveling the mechanisms of synapse formation and axon Shao, Z., Watanabe, S., Christensen, R., Jorgensen, E. M., and Colon-Ramos, D.
regeneration: the awesome power of C. elegans genetics. Sci. China Life Sci. 58, A. (2013). Synapse location during growth depends on glia location. Cell 154,
1084–1088. doi: 10.1007/s11427-015-4962-9 337–350. doi: 10.1016/j.cell.2013.06.028
Kim, J., Zhao, T., Petralia, R. S., Yu, Y., Peng, H., Myers, E., et al. (2012). mGRASP Shen, K., and Bargmann, C. I. (2003). The immunoglobulin superfamily protein
enables mapping mammalian synaptic connectivity with light microscopy. Nat. SYG-1 determines the location of specific synapses in C. elegans. Cell 112,
Methods 9, 96–102. doi: 10.1038/nmeth.1784 619–630. doi: 10.1016/S0092-8674(03)00113-2
Kittelmann, M., Hegermann, J., Goncharov, A., Taru, H., Ellisman, M. H., Shen, K., Fetter, R. D., and Bargmann, C. I. (2004). Synaptic specificity is generated
Richmond, J. E., et al. (2013). Liprin-alpha/SYD-2 determines the size of dense by the synaptic guidepost protein SYG-2 and its receptor, SYG-1. Cell 116,
projections in presynaptic active zones in C. elegans. J. Cell Biol. 203, 849–863. 869–881. doi: 10.1016/S0092-8674(04)00251-X
doi: 10.1083/jcb.201302022 Shipley, F. B., Clark, C. M., Alkema, M. J., and Leifer, A. M. (2014). Simultaneous
Kocabas, A., Shen, C. H., Guo, Z. V., and Ramanathan, S. (2012). Controlling optogenetic manipulation and calcium imaging in freely moving C. elegans.
interneuron activity in Caenorhabditis elegans to evoke chemotactic behaviour. Front. Neural Circuits 8:28. doi: 10.3389/fncir.2014.00028
Nature 490, 273–277. doi: 10.1038/nature11431 Sieburth, D., Ch’ng, Q., Dybbs, M., Tavazoie, M., Kennedy, S., Wang, D.,
Kuner, T., and Augustine, G. J. (2000). A genetically encoded ratiometric indicator et al. (2005). Systematic analysis of genes required for synapse structure and
for chloride: capturing chloride transients in cultured hippocampal neurons. function. Nature 436, 510–517. doi: 10.1038/nature03809
Neuron 27, 447–459. doi: 10.1016/S0896-6273(00)00056-8 Stigloher, C., Zhan, H., Zhen, M., Richmond, J., and Bessereau, J. L. (2011).
Leifer, A. M., Fang-Yen, C., Gershow, M., Alkema, M. J., and Samuel, A. D. (2011). The presynaptic dense projection of the Caenorhabditis elegans cholinergic
Optogenetic manipulation of neural activity in freely moving Caenorhabditis neuromuscular junction localizes synaptic vesicles at the active zone through
elegans. Nat. Methods 8, 147–152. doi: 10.1038/nmeth.1554 SYD-2/liprin and UNC-10/RIM-dependent interactions. J. Neurosci. 31,
Li, Y., Guo, A., and Li, H. (2016). CRASP: CFP reconstitution across 4388–4396. doi: 10.1523/JNEUROSCI.6164-10.2011
synaptic partners. Biochem. Biophys. Res. Commun. 469, 352–356. doi: Stirman, J. N., Crane, M. M., Husson, S. J., Wabnig, S., Schultheis, C., Gottschalk,
10.1016/j.bbrc.2015.12.011 A., et al. (2011). Real-time multimodal optical control of neurons and muscles
Luo, L., Wen, Q., Ren, J., Hendricks, M., Gershow, M., Qin, Y., et al. (2014). in freely behaving Caenorhabditis elegans. Nat. Methods 8, 153–158. doi:
Dynamic encoding of perception, memory, and movement in a C. elegans 10.1038/nmeth.1555
chemotaxis circuit. Neuron 82, 1115–1128. doi: 10.1016/j.neuron.2014.05.010 Sulston, J. E., and White, J. G. (1980). Regulation and cell autonomy during
Macpherson, L. J., Zaharieva, E. E., Kearney, P. J., Alpert, M. H., Lin, T. Y., Turan, postembryonic development of Caenorhabditis elegans. Dev. Biol. 78, 577–597.
Z., et al. (2015). Dynamic labelling of neural connections in multiple colours doi: 10.1016/0012-1606(80)90353-X
by trans-synaptic fluorescence complementation. Nat. Commun. 6:10024. doi: Tian, L., Hires, S. A., Mao, T., Huber, D., Chiappe, M. E., Chalasani, S. H., et al.
10.1038/ncomms10024 (2009). Imaging neural activity in worms, flies and mice with improved GCaMP
Margeta, M. A., and Shen, K. (2010). Molecular mechanisms of synaptic specificity. calcium indicators. Nat. Methods 6, 875–881. doi: 10.1038/nmeth.1398
Mol. Cell. Neurosci. 43, 261–267. doi: 10.1016/j.mcn.2009.11.009 Trojanowski, N. F., Padovan-Merhar, O., Raizen, D. M., and Fang-Yen, C.
McIntire, S. L., Jorgensen, E., Kaplan, J., and Horvitz, H. R. (1993). The (2014). Neural and genetic degeneracy underlies Caenorhabditis elegans feeding
GABAergic nervous system of Caenorhabditis elegans. Nature 364, 337–341. behavior. J. Neurophysiol. 112, 951–961. doi: 10.1152/jn.00150.2014
doi: 10.1038/364337a0 Tsetsenis, T., Boucard, A. A., Arac, D., Brunger, A. T., and Sudhof,
Mello, C. C., Kramer, J. M., Stinchcomb, D., and Ambros, V. (1991). Efficient T. C. (2014). Direct visualization of trans-synaptic neurexin-neuroligin
gene transfer in C.elegans: extrachromosomal maintenance and integration of interactions during synapse formation. J. Neurosci. 34, 15083–15096. doi:
transforming sequences. EMBO J. 10, 3959–3970. 10.1523/JNEUROSCI.0348-14.2014
Nagel, G., Brauner, M., Liewald, J. F., Adeishvili, N., Bamberg, E., and Gottschalk, Umemori, H., Linhoff, M. W., Ornitz, D. M., and Sanes, J. R. (2004). FGF22 and its
A. (2005). Light activation of channelrhodopsin-2 in excitable cells of close relatives are presynaptic organizing molecules in the mammalian brain.
Caenorhabditis elegans triggers rapid behavioral responses. Curr. Biol. 15, Cell 118, 257–270. doi: 10.1016/j.cell.2004.06.025
2279–2284. doi: 10.1016/j.cub.2005.11.032 Wadsworth, W. G., Bhatt, H., and Hedgecock, E. M. (1996). Neuroglia and pioneer
Nagel, G., Szellas, T., Huhn, W., Kateriya, S., Adeishvili, N., Berthold, P., neurons express UNC-6 to provide global and local netrin cues for guiding
et al. (2003). Channelrhodopsin-2, a directly light-gated cation-selective migrations in C. elegans. Neuron 16, 35–46. doi: 10.1016/S0896-6273(00)80021-
membrane channel. Proc. Natl. Acad. Sci. U.S.A. 100, 13940–13945. doi: 5
10.1073/pnas.1936192100 Waschuk, S. A., Bezerra, A. G. Jr., Shi, L., and Brown, L. S. (2005). Leptosphaeria
Nonet, M. L. (1999). Visualization of synaptic specializations in live C. elegans rhodopsin: bacteriorhodopsin-like proton pump from a eukaryote. Proc. Natl.
with synaptic vesicle protein-GFP fusions. J. Neurosci. Methods 89, 33–40. doi: Acad. Sci. U.S.A. 102, 6879–6883. doi: 10.1073/pnas.0409659102
10.1016/S0165-0270(99)00031-X Watanabe, S., Punge, A., Hollopeter, G., Willig, K. I., Hobson, R. J., Davis, M. W.,
Park, M., and Shen, K. (2012). WNTs in synapse formation and neuronal circuitry. et al. (2011). Protein localization in electron micrographs using fluorescence
EMBO J. 31, 2697–2704. doi: 10.1038/emboj.2012.145 nanoscopy. Nat. Methods 8, 80–84. doi: 10.1038/nmeth.1537
Patel, M. R., Lehrman, E. K., Poon, V. Y., Crump, J. G., Zhen, M., Bargmann, C. Weimer, R. M., Gracheva, E. O., Meyrignac, O., Miller, K. G., Richmond, J.
I., et al. (2006). Hierarchical assembly of presynaptic components in defined C. E., and Bessereau, J. L. (2006). UNC-13 and UNC-10/rim localize synaptic
elegans synapses. Nat. Neurosci. 9, 1488–1498. doi: 10.1038/nn1806 vesicles to specific membrane domains. J. Neurosci. 26, 8040–8047. doi:
Patel, M. R., and Shen, K. (2009). RSY-1 is a local inhibitor of presynaptic assembly 10.1523/JNEUROSCI.2350-06.2006
in C. elegans. Science 323, 1500–1503. doi: 10.1126/science.1169025 White, E. L. (2007). Reflections on the specificity of synaptic connections. Brain
Rabinowitch, I., Chatzigeorgiou, M., and Schafer, W. R. (2013). A gap junction Res. Rev. 55, 422–429. doi: 10.1016/j.brainresrev.2006.12.004
circuit enhances processing of coincident mechanosensory inputs. Curr. Biol. White, J. G., Southgate, E., Thomson, J. N., and Brenner, S. (1986). The structure
23, 963–967. doi: 10.1016/j.cub.2013.04.030 of the nervous system of the nematode Caenorhabditis elegans. Philos. Trans. R.
Rabinowitch, I., Chatzigeorgiou, M., Zhao, B., Treinin, M., and Schafer, W. R. Soc. Lond. B. Biol. Sci. 314, 1–340. doi: 10.1098/rstb.1986.0056
(2014). Rewiring neural circuits by the insertion of ectopic electrical synapses Wu, Y., Ghitani, A., Christensen, R., Santella, A., Du, Z., Rondeau, G., et al. (2011).
in transgenic C. elegans. Nat. Commun. 5:4442. doi: 10.1038/ncomms5442 Inverted selective plane illumination microscopy (iSPIM) enables coupled cell
identity lineaging and neurodevelopmental imaging in Caenorhabditis elegans. Zhang, F., Prigge, M., Beyriere, F., Tsunoda, S. P., Mattis, J., Yizhar, O., et al. (2008).
Proc. Natl. Acad. Sci. U.S.A. 108, 17708–17713. doi: 10.1073/pnas.1108494108 Red-shifted optogenetic excitation: a tool for fast neural control derived from
Yamagata, M., and Sanes, J. R. (2012). Transgenic strategy for identifying synaptic Volvox carteri. Nat. Neurosci. 11, 631–633. doi: 10.1038/nn.2120
connections in mice by fluorescence complementation (GRASP). Front. Mol. Zhang, F., Wang, L. P., Brauner, M., Liewald, J. F., Kay, K., Watzke, N., et al. (2007).
Neurosci. 5:18. doi: 10.3389/fnmol.2012.00018 Multimodal fast optical interrogation of neural circuitry. Nature 446, 633–639.
Yamagata, M., Weiner, J. A., and Sanes, J. R. (2002). Sidekicks: synaptic adhesion doi: 10.1038/nature05744
molecules that promote lamina-specific connectivity in the retina. Cell 110, Zhen, M., and Jin, Y. (1999). The liprin protein SYD-2 regulates the differentiation
649–660. doi: 10.1016/S0092-8674(02)00910-8 of presynaptic termini in C. elegans. Nature 401, 371–375. doi: 10.1038/43886
Yeh, E., Kawano, T., Weimer, R. M., Bessereau, J. L., and Zhen, M. (2005). Zhen, M., and Samuel, A. D. (2015). C. elegans locomotion: small circuits, complex
Identification of genes involved in synaptogenesis using a fluorescent active functions. Curr. Opin. Neurobiol. 33, 117–126. doi: 10.1016/j.conb.2015.
zone marker in Caenorhabditis elegans. J. Neurosci. 25, 3833–3841. doi: 03.009
10.1523/JNEUROSCI.4978-04.2005
Yogev, S., and Shen, K. (2014). Cellular and molecular mechanisms of synaptic Conflict of Interest Statement: The authors declare that the research was
specificity. Annu. Rev. Cell Dev. Biol. 30, 417–437. doi: 10.1146/annurev- conducted in the absence of any commercial or financial relationships that could
cellbio-100913-012953 be construed as a potential conflict of interest.
Zhai, R. G., and Bellen, H. J. (2004). The architecture of the active zone
in the presynaptic nerve terminal. Physiology (Bethesda) 19, 262–270. doi: Copyright © 2016 Hong and Park. This is an open-access article distributed under the
10.1152/physiol.00014.2004 terms of the Creative Commons Attribution License (CC BY). The use, distribution or
Zhan, H., Stanciauskas, R., Stigloher, C., Dizon, K. K., Jospin, M., Bessereau, J. reproduction in other forums is permitted, provided the original author(s) or licensor
L., et al. (2014). In vivo single-molecule imaging identifies altered dynamics of are credited and that the original publication in this journal is cited, in accordance
calcium channels in dystrophin-mutant C. elegans. Nat. Commun. 5, 4974. doi: with accepted academic practice. No use, distribution or reproduction is permitted
10.1038/ncomms5974 which does not comply with these terms.