Abstract
Neuroscientists spent decades debating whether synaptic plasticity was presynaptically or postsynaptically expressed. It was eventually concluded that plasticity depends on many factors, including cell type. More recently, it has become increasingly clear that plasticity is regulated at an even finer grained level; it is specific to the synapse type, a concept we denote synapse-type-specific plasticity (STSP). Here, we review recent developments in the field of STSP, discussing both long-term and short-term variants and with particular emphasis on neocortical function. As there are dozens of neocortical cell types, there is a multiplicity of forms of STSP, the vast majority of which have never been explored. We argue that to understand the brain and synaptic diseases, we have to grapple with STSP.
Introduction
Neuroscientists spent decades debating whether synaptic plasticity was presynaptically or postsynaptically expressed. It was eventually concluded that plasticity depends on many factors, including cell type. More recently, it has become increasingly clear that plasticity is regulated at an even finer grained level; it is specific to the synapse type, a concept we denote synapse-type-specific plasticity (STSP). Here, we review recent developments in the field of STSP, discussing both long-term and short-term variants and with particular emphasis on neocortical function. As there are dozens of neocortical cell types, there is a multiplicity of forms of STSP, the vast majority of which have never been explored. We argue that to understand the brain and synaptic diseases, we have to grapple with STSP.
A definition of synapse-type-specific plasticity
Here, we define synapse-type-specific plasticity (STSP) as plasticity that varies with synapse type. STSP comes in different forms (Figure 1). Different synapses originating from the same presynaptic cell can have different forms of plasticity depending upon the on postsynaptic target, here termed divergent STSP. Conversely, plasticity may differ for the same postsynaptic cell depending on the type of input, which we denote convergent STSP. While more difficult to distinguish experimentally, STSP also encompasses plasticity that varies at the level of individual synaptic contacts between one presynaptic axon and individual postsynaptic neuron. In this review, we discuss examples of all three forms of STSP.
Figure 1.
Different forms of synapse-type-specific plasticity. Synapse-type-specific plasticity (STSP) comes in several different forms. It may be divergent, in which case synapses from the presynaptic cell type A to the postsynaptic cell type B do not undergo the same kind of plasticity as synapses onto type C. An example of this is illustrated in Figure 2 [9••]. STSP may also be convergent, so that connections from presynaptic cell type A onto postsynaptic cell type C do not have the same kind of plasticity as connections from presynaptic cell type B onto C do. This scenario is illustrated in Figure 3 [31••,67••]. Many connections in the brain, however, are made from multiple synaptic contacts [5], and a third possible STSP scenario exists in which plasticity is contact specific. This scenario is less well studied, but some studies indicate that different contacts of the same connection type are very similar with respect to plasticity [5,62•], whereas others imply that they may be different [52•,60].
STSP should not be mixed up with synapse or input specificity in classical plasticity studies. A form of long-term plasticity that is input specific does not spread appreciably to neighboring synapses that were not stimulated during the induction of plasticity [1]. In other words, STSP may or may not be input specific. In the case of long-term plasticity, STSP is therefore not directly related to either homosynaptic or heterosynaptic plasticity [1].
Synaptic plasticity may vary in several different ways. STSP may thus refer to distinct specificities in terms of phenomenology of plasticity, induction mechanism, and/ or expression mechanism. Finally, STSP may also refer to either short or long-term plasticity, and we review both these scenarios here below, starting with the former.
Synapse-type-specific short-term plasticity
Short-term plasticity (STP) refers to a depression or facilitation of synaptic efficacy that last on the order of seconds [2]. The mechanisms underlying STP are typically presynaptic, including changes in the readily releasable vesicle pool size, changes in the number of release sites, and alterations in presynaptic calcium dynamics [3]. Early evidence that STP could be synapse-type-specific came from observations that the magnitude and variability of calcium influx in response to single action potentials was not uniform across a single axon collateral [4,5]. In addition, these differences in calcium influx across a single axon have been shown to contribute to distinct vesicular release probabilities at individual synapses [6,7]. Synapse-type-specific STP thought to result from such differences has been observed in both neocortex and hippocampus [8,9••,10]. These differences in STP are also likely correlated with differences in synaptic morphology. For example, the number of docked vesicles is known to be correlated with both presynaptic release probability and the readily releasable synaptic vesicle pool [11,12]. However, differences in docked vesicles do not account for different STP at different presynaptic terminals onto the same postsynaptic cell [13], suggesting these ultrastructural measurements may not always be sufficient to infer synapse-type-specific STP alone.
STP differs with both pre-synaptic and post-synaptic cell type, suggesting that specific trans-synaptic signaling mechanisms define STP at a given synapse. These mechanisms are poorly understood, but recent findings suggest that extracellular leucine-rich repeat proteins such as Elfn1 trans-synaptically regulate synapse-type-specific STP [14••]. In hippocampus, Elfn1 is selectively expressed at somatostatin but not parvalbumin-positive interneurons, where it signals retrogradely to increase facilitation at glutamatergic inputs. Adhesion molecules such as cadherins, neuroligins, or SynGAP also regulate synaptic dynamics [15•,16,17,18]. In addition, recent studies suggest that synapse-type-specific STP is influenced presynaptically by RIM, α2δ, calcineurin Aα, and CDK5, proteins that regulate release probability and STP by altering presynaptic calcium channel expression [19–21]. Similarly, synapse-type-specific expression of voltage-gated Na+ and K+ channels at individual axon boutons may influence action potential waveforms distinctly at individual synapses [22,23•]. A general mechanism for the specification of synapse-type-specific STP may therefore be cell-type specific expression of Elfn1 or related molecules, which govern the synapse-type-specific expression of adhesion molecules. These in turn may change the expression of voltage-gated ion channels or presynaptic neurotransmitter receptors, ultimately altering vesicular release [24].
Genetic programs are not the sole determinants of STP, as it depends on age and activity. During development, for example, STP generally becomes less depressing [25–30]. This developmental change, however, is not homogenous across all synapse types. Connections between cortical layer 5 (L5) and L2/3 pyramidal neurons undergo a reduction in synaptic depression through development, but the reduction is larger at synapses between L5 neurons [25]. At L2/3 visual cortical neurons, L4 synapses undergo a reduction in short-term depression through development, but at intralaminar L2/3 synapses onto the same postsynaptic cell type do not undergo this change [27,31••,32••,33]. These layer-specific differences in STP may result from different developmental time points for layer formation [34], and therefore these differences may not always last into adulthood.
Developmental alterations in STP may involve both pre-synaptic and post-synaptic changes, demonstrating that synapse-type-specificity is not universally dictated by a single mechanism. For example, short-range but not long-range excitatory inputs onto parvalbumin-positive interneurons of visual cortex undergo a developmental reduction in short-term depression involving both a reduction in presynaptic release and increase in calcium-permeable AMPARs [32••]. In this case, increasing the number of calcium-permeable AMPARs increases short-term facilitation postsynaptically through a reduction in polyamine-dependent calcium-permeable AMPAR block during repetitive presynaptic cell firing [35,36]. Polyamine-dependent facilitation of AMPARs [35,36] and other postsynaptic phenomenology such as temporal summation [37] may be general mechanisms for postsynaptically mediated short-term STSP. In agreement with this view, CP-AMPARs are only expressed at specific subsets of synapses [36]. Overall, these findings suggest there is a developmental synapse-type-specific alteration in STP, rather than a uniform reduction in synaptic depression across all synapses.
Sensory experience also alters STP, however it is unclear whether experience per se is required because such alterations are also observed in cultured neurons [38]; perhaps neuronal activity is sufficient to drive changes in STP. Conversely, sensory deprivation alters STP at a variety of synapses. However, the effect of sensory deprivation on STP is often inconsistent even for the same synapse type, with specific types of sensory deprivation promoting facilitation but others increasing depression [31••,39–43]. A number of experimental factors, including the nature and age of deprivation [39,44], rate of presynaptic stimulation or firing [31••], and pharmacological receptor blockade [31••] likely contribute to such inconsistencies. These developmental alterations in STP are probably key to regulating neuronal excitation because excitatory short-term synaptic transmission depresses more when inhibition is still immature [27]. Subsequently, experience may alter STP to enable reliable information processing and to promote Hebbian plasticity.
Recent evidence suggests that synapse-type-specific expression of presynaptic NMDA receptors (preNMDARs) is an important determinant of developmental and sensory-experience driven alterations in STP. Within visual cortical L5, preNMDARs selectively influence STP at pyramidal cells synapsing onto neighboring pyramidal cells or onto somatostatin-positive Martinotti cells, but not onto parvalbumin-positive basket cells [9••] (Figure 2a). Pre-NMDARs typically modulate STP by acting as a high-pass filter for presynaptic release [31••,45,46]. Through this action, preNMDARs at excitatory synapses onto L5 Martinotti cells increase frequency-dependent disynaptic inhibition (Figure 2b–d). Similarly, preNMDARs influence STP at synapses from L4 neurons onto L2/3 pyramidal neurons, but not at L2/3 intralaminar synapses [47]. Development and sensory experience regulate pre-NMDAR expression, suggesting that these receptors are well suited for regulating synapse-type-specific STP [31••,33,39,48]. Synapse-type-specific expression of pre-NMDARs may thus be a general principle governing the function of this receptor type. Since preNMDAR expression varies with postsynaptic cell identity, we propose that postsynaptic neurons retrogradely influence the expression of preNMDARs at specific synapse types. This may occur through postsynaptic proteins such as Elfn1, which is known to influence presynaptic ionotropic receptor expression [14••].
Figure 2.
PreNMDARs synapse-type-specifically reroute information flow in local neocortical circuits. (a) PreNMDARs are synapse-type-specifically expressed at connections (closed symbols) from pyramidal cells (‘PC’) to other pyramidal cells as well as to Martinotti cells (‘MC’), but not at connections (open symbols) to basket cells (‘BC’). Circles and triangles symbolize inhibitory and excitatory connections, respectively. (b) A small network model with synaptic dynamics tuned to experimental data predicted that during 70-Hz pyramidal-cell bursts (vertical lines), preNMDAR blockade would disrupt late Martinotti but not early basket-cell inhibition. Model prediction before and after preNMDAR blockade in blue and red, respectively. (c) 70-Hz spiking in pyramidal cell 1 evoked both Martinotti and basket-cell inhibition in pyramidal cell 3. Intermediate Martinotti cell (‘X’) was not patched. (d) Experiments verified model prediction: amplitude and latency of Martinotti but not basket-cell inhibition was affected by preNMDAR blockade. In other words, synapse-type-specific preNMDAR expression selectively boosts pyramid to Martinotti cell neurotransmission during high-frequency firing without affecting pyramidal cell connections to basket cells. Modified with permission from [9••].
Synapse-type-specific long-term plasticity
Long-term plasticity refers to activity-dependent changes in synaptic efficacy that last for minutes up to days [49]. Long-term plasticity is typically mechanistically distinct from STP and includes processes such as Hebbian and homeostatic plasticity. Long-lasting increases or decreases in synaptic strength — long-term potentiation (LTP) or depression (LTD), respectively — are believed to be critical to circuit and memory formation [50]. Synapse-type-specific long-term plasticity may thus be important for adapting synapses to the overall function of the circuit in which they are embedded.
Homosynaptic forms of long-term plasticity are specific to activated synapses. The underlying mechanisms can be synapse-type-specific as well, resulting in differences in locus of expression or in activity requirements. Classically, LTP and LTD are induced by the activation of postsynaptic NMDA, AMPA, and metabotropic glutamate receptors (mGluRs). The expression of these postsynaptic receptors as well as their presynaptic counterparts varies considerably with brain region, synapse type, and through development [51]. At large mature dendritic spines, LTD requires activation of both NMDARs and mGluRs as well as calcium release from internal stores [52•,53]. However, at small immature spines, LTD requires only NMDAR activation [52•,53]. Since large spines contain more AMPARs [54], these findings are consistent with synapse-type-specific metaplasticity, in which previous history of activity alters the subsequent threshold for plasticity [55]. Indeed, sensory experience and neuronal activity alter postsynaptic NMDAR subunit expression at many synapses within the brain [56]. In hippocampal culture, alterations in synaptic activity at single synapses are sufficient to alter both postsynaptic NMDAR expression and later outcomes of plasticity, demonstrating that metaplasticity is an important determinant for synapse-type-specific long-term plasticity [57]. Synapse-type-specific differences in the expression of proteins signaling cascades downstream of neurotransmitter receptors such as CaM-KII and MAPK may also be important for synapse-type-specific expression of plasticity, but how the expression of these molecules differs for a given neuron type based on synaptic source is largely unknown [58]. However, differences in postsynaptic signaling can influence multiple forms of long-term plasticity since synapse-type-specific expression of activity related proteins such as Arc have been shown to influence synapse-type-specific homeostatic plasticity [59]. The initial properties of long-term synaptic plasticity may thus be broadly defined in synapse-type-specific manners. Through alterations in neuronal activity, synapses may subsequently acquire specific signaling molecules that enable synapse-type-specific long-term plasticity.
Dendrite biophysics may also help determine long-term STSP. For example, distal excitatory synapses onto neo-cortical L5 pyramidal cells are less prone to long-term potentiation than the proximal ones are [60], because backpropagating action potentials — which determine potentiation in this cell type — fail to invade distal arbors [1]. But each synaptic connection is typically formed by multiple synaptic contacts, some quite distal and some proximal [61], which implies that plasticity might be synaptic contact specific, since distal contacts may undergo LTD when proximal contacts potentiate. Interestingly, in the hippocampal CA1 area, pairs of spines on the same dendrite that received input from the same axon are of similar size, whereas spines on different dendrites are not [62•,63]. Because spine volume is closely correlated with synaptic strength [12], this is concrete evidence that plasticity is specific even down to individual synaptic contacts [5].
PreNMDARs also mediate diverse forms of LTP and LTD at subsets of synapses, including timing-dependent LTD at cortical synapses and theta-burst stimulation induced cortical-striatal LTP [45,64–66]. PreNMDARs are therefore likely to bias plasticity and circuit modifica-tion. In early neocortical development, preNMDARs mediate timing-dependent LTD (tLTD) at L4 to L2/3 synapses [48,66], but tLTD at horizontal L2/3 to L2/3 synapses requires postsynaptic NMDARs [67••,68]. These mechanistic differences may underlie the considerably longer tLTD window at L4 to L2/3 synapses [67••] (Figure 3a–c). Due to developmental downregulation of preNMDARs, tLTD at L4-to-L2/3, but not at L2/3-to-L2/ 3 synapses, is more readily induced in early development and becomes increasingly dependent on postsynaptic NMDAR activation in later development [48,69]. In the visual cortex, dark rearing prolongs preNMDAR-mediated tLTD at L4 to L2/3 synapses, suggesting preNMDARs enable specific synapses to uniquely respond to changes in the sensory environment [31••] (Figure 3d–e). Taken together, these studies show that preNMDARs are well positioned for synapse-type-specific control of both long and short-term plasticity.
Figure 3.
Experience-dependent regulation of synapse-type-specific long-term depression at layer 2/3 cortical synapses. In early cortical development, timing-dependent LTD (tLTD) at L4 to L2/3 synapses requires preNMDARs and can be induced with relatively long action potential-EPSP pairings (a, c). In contrast, tLTD at L2/3 to L2/3 horizontal connections requires postsynaptic NMDARs and has a shorter tLTD integration window (b, c). PreNMDAR-mediated tLTD at L4 to L2/3 is developmentally downregulated following the first 3–4 weeks of postnatal development, but this downregulation does not occur at visual cortical L2/3 to L2/3 synapses [31••,69]. Thus, in normally reared P85–95 mice, pairing action potentials with EPSPs generated by optogenetic stimulation of L4 neurons fails to induce tLTD at L4 to L2/3 synapses (d–e, yellow circles). However, visually depriving mice for 10–15 days during this developmental period restores preNMDAR-mediated tLTD at L4 to L2/3 synapses (e, gray circles). This effect of visual-deprivation is lost when preNMDARs are selectively deleted from L4 neurons (e, orange circles). Modified with permission from [31••,67••].
Conclusions and future directions
In this review, we have discussed recent research on STSP learning rules and synaptic dynamics, as well as their potential functional roles (summarized in Table 1). We argue that STSP may have evolved of necessity to enable more complex computations in local circuits. The existence of STSP is furthermore expected, since the nodes of biological neuronal networks are made up of dozens of different cell types [70–72]. Presumably, these different cell types need to be governed by distinct synaptic plasticity learning rules, because activity patterns and excitabilities vary tremendously with cell type. Indeed, it was recently shown that pattern separation in the cerebellum is enhanced by input-specific STP onto granule cells, since the distinct activity patterns of modality-specific afferents means coding benefits from differential short-term STSP [73••]. Furthermore, activity levels must be kept within reasonable bounds, and it is likely that one or several interneuron types specifically adjust their inhibitory influence according to key developmental events such as eye opening, or in pathological states such as physical trauma, inflammation, or epilepsy. Indeed, STSP dysfunction has been observed in a variety of disease states ranging from schizophrenia [74], Alzheimer’s disease [75], to autism spectrum disorders [76,77], demonstrating that an understanding of the properties of STSP may be fundamental to treating neurological diseases.
Table 1.
Examples of STSP.
| Brain region | Synapse types | Consequences | Mechanisms | Ref |
|---|---|---|---|---|
| Visual cortex | Excitatory connections from L4 neurons to L2/3 PCs | Experience-dependent regulation of LTD and STP | PreNMDARs | [31••] |
| Barrel cortex | Excitatory connections from L4 neurons to L2/3 PCs | Experience-dependent regulation of STP | PreNMDARs | [39] |
| Barrel cortex | Excitatory connections from L4 neurons to L2/3 PCs | Trigger LTD | PreNMDARs | [67••] |
| Barrel cortex | Excitatory connections from L4 neurons to L2/3 PCs | Upregulate short-term depression | PreNMDARs | [47] |
| Visual cortex | Excitatory connections between L5 PCs | Evoke LTD and upregulate short- term depression | PreNMDARs | [45] |
| Visual cortex | Excitatory connections from L5 PCs to MCs | Upregulate dendritic inhibition | PreNMDARs | [9••] |
| Visual Cortex | Short-range excitatory connections from PCs to L2/3 PV-positive basket cells | Increase short-term facilitation through polyamine-dependent AMPAR facilitation | CP-AMPARs | [32••] |
| Hippocampus | Excitatory connections from CA1 PCs to Sst-positive INs | Upregulate short-term facilitation | Presynaptic kainate receptors, postsynaptic Elfn1 | [14••,78] |
| Hippocampus | Excitatory contacts to CA1 PCs | Homogenous STP for synaptic contacts with the same presynaptic and postsynaptic cells | Synaptic plasticity | [62•,63] |
| Hippocampus | Hippocampal CA3 and CA1 dissociated culture | Regulate presynaptic release probability through modulation of calcium channel number and/or function | VGCC auxiliary subunit α2δ, calcineurin Aα, CDK5 | [19,20] |
| Hippocampus | Large excitatory spines onto CA1 PCs | Regulates spine shrinkage, LTD | Type 1 mGluRs, calcium release from internal stores | [52•] |
| Cerebellum | Purkinje and basket cell terminals in dissociated culture | Regulation of STP by control of axon terminal excitability | Differential Na+/K+ channel ratio | [23•] |
| Cerebellum | Excitatory inputs to granule cells | Facilitation of pattern separation | Modality-specific STP | [73••] |
The studies in this table illustrate in several different brain regions how STSP can be produced mechanistically and what the functional consequences can be. This list is not intended to be exhaustive.
While there are many mechanisms for regulating STSP, we have focused on the preNMDAR as one important example of a determinant of STSP in local circuits that controls both short-term [9••] and long-term plasticity [67••] and that is linked to the closing of the visual cortex critical period [31••]. This preNMDAR focus is not to imply that this receptor type necessarily holds a special status in the brain; it is merely a case study and a starting point for further research on STSP. Indeed, we suggest that control of STSP by particular receptor types is a general principle in the brain. For example, presynaptic kainate receptors also determine STSP, by promoting facilitation at excitatory inputs onto somatostatin-positive hippocampal interneurons [14••,15•,78].
When it comes to complexity, neuronal network computer models are invariably a far cry from their biological counterparts. In models, synaptic weights typically have no short-term dynamics, neurons do not accommodate, that is if they spike at all, and nodes are often perfectly homogenous. This is natural, as most computer models are not intended to simulate detail, nor do they need to. But for aficionados of detail, this is of course entirely wrong. Large-scale detailed computer models of neuronal circuits have recently received much attention as well as criticism [79,80]. Although a natural next step in computer modeling might seem to require incorporating considerably more detail, our review highlights just how much detail there is. For every cell type — and there is a lot of them [70–72] — there is a corresponding multiplicity of STSP learning rules and short-term dynamics. Yet, to understand the brain and synaptic diseases such as autism, anxiety, or epilepsy [81], it appears that we have no choice but to grapple with STSP. The task that lies ahead is nothing short of formidable, as we have just barely scratched the surface.
Acknowledgments
We thank Alanna Watt, Ben Philpot, Ole Paulsen, Abhishek Banerjee, Txomin Lalanne, Elvis Cela, Therese Abrahamsson, Arne Blackman, and Rui P. Costa for help and useful discussions. R.S.L. was funded by the Allen Institute for Brain Science and wishes to thank the Allen Institute founders, Paul G. Allen and Jody Allen, for their vision, encouragement, and support. P.J.S. was funded by CFI LOF 28331, CIHR OG 126137, CIHR NIA 288936, and NSERC DG 418546-2.
References and recommended reading
Papers of particular interest, published within the period of review, have been highlighted as:
• of special interest
•• of outstanding interest
- 1.Sjöström PJ, Rancz EA, Roth A, Häusser M. Dendritic excitability and synaptic plasticity. Physiol Rev. 2008;88:769–840. doi: 10.1152/physrev.00016.2007. [DOI] [PubMed] [Google Scholar]
- 2.Stevens CF, Wang Y. Facilitation and depression at single central synapses. Neuron. 1995;14:795–802. doi: 10.1016/0896-6273(95)90223-6. [DOI] [PubMed] [Google Scholar]
- 3.Regehr WG. Short-term presynaptic plasticity. Cold Spring Harb Perspect Biol. 2012;4:a005702. doi: 10.1101/cshperspect.a005702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Frenguelli BG, Malinow R. Fluctuations in intracellular calcium responses to action potentials in single en passage presynaptic boutons of layer V neurons in neocortical slices. Learn Mem. 1996;3:150–159. doi: 10.1101/lm.3.2-3.150. [DOI] [PubMed] [Google Scholar]
- 5.Koester HJ, Johnston D. Target cell-dependent normalization of transmitter release at neocortical synapses. Science. 2005;308:863–866. doi: 10.1126/science.1100815. [DOI] [PubMed] [Google Scholar]
- 6.Ermolyuk YS, Alder FG, Henneberger C, Rusakov DA, Kullmann DM, Volynski KE. Independent regulation of basal neurotransmitter release efficacy by variable Ca(2)+ influx and bouton size at small central synapses. PLoS Biol. 2012;10:e1001396. doi: 10.1371/journal.pbio.1001396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Ariel P, Hoppa MB, Ryan TA. Intrinsic variability in Pv, RRP size, Ca(2+) channel repertoire, and presynaptic potentiation in individual synaptic boutons. Front Synaptic Neurosci. 2012;4:9. doi: 10.3389/fnsyn.2012.00009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Ali AB, Deuchars J, Pawelzik H, Thomson AM. CA1 pyramidal to basket and bistratified cell EPSPs: dual intracellular recordings in rat hippocampal slices. J Physiol. 1998;507(Pt 1):201–217. doi: 10.1111/j.1469-7793.1998.201bu.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9••.Buchanan KA, Blackman AV, Moreau AW, Elgar D, Costa RP, Lalanne T, Tudor Jones AA, Oyrer J, Sjöström PJ. Target-specific expression of presynaptic NMDA receptors in neocortical microcircuits. Neuron. 2012;75:451–466. doi: 10.1016/j.neuron.2012.06.017. This study uses 2-photon calcium imaging, NMDA uncaging, computer modeling, and paired-recordings between L5 neurons to evaluate the influence of preNMDARs on STP at excitatory-to-excitatory and excitatory-to-inhibitory synapses. The authors find that preNMDARs influence both STP and presynaptic calcium signaling at a specific subset of axonal boutons, demonstrating that these receptors underlie STSP of synaptic dynamics. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Markram H, Wang Y, Tsodyks M. Differential signaling via the same axon of neocortical pyramidal neurons. Proc Natl Acad Sci U S A. 1998;95:5323–5328. doi: 10.1073/pnas.95.9.5323. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Branco T, Marra V, Staras K. Examining size–strength relationships at hippocampal synapses using an ultrastructural measurement of synaptic release probability. J Struct Biol. 2010;172:203–210. doi: 10.1016/j.jsb.2009.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Murthy VN, Schikorski T, Stevens CF, Zhu Y. Inactivity produces increases in neurotransmitter release and synapse size. Neuron. 2001;32:673–682. doi: 10.1016/s0896-6273(01)00500-1. [DOI] [PubMed] [Google Scholar]
- 13.Xu-Friedman MA, Harris KM, Regehr WG. Three-dimensional comparison of ultrastructural characteristics at depressing and facilitating synapses onto cerebellar Purkinje cells. J Neurosci. 2001;21:6666–6672. doi: 10.1523/JNEUROSCI.21-17-06666.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14••.Sylwestrak EL, Ghosh A. Elfn1 regulates target-specific release probability at CA1-interneuron synapses. Science. 2012;338:536–540. doi: 10.1126/science.1222482. Through analysis of hippocampal mRNA expression, the authors identify the leucine-rich repeat protein Elfn1 that is expressed in a group of somatostatin-positive neurons and which increases synaptic facilitation at glutamatergic synapses onto these neurons. When exogenously expressed in parvalbumin-positive neurons, Elfn1 transforms their normally short-term depressing glutamatergic inputs into facilitating ones resembling those of somatostatin-positive neurons, demonstrating Elfn1’s involvement in synapse-type-specific retrograde regulation of STP. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15•.Blackman AV, Abrahamsson T, Costa RP, Lalanne T, Sjöström PJ. Target-cell-specific short-term plasticity in local circuits. Front Synaptic Neurosci. 2013;5:11. doi: 10.3389/fnsyn.2013.00011. The authors provide a detailed review of short-term STSP. They suggest specific functional roles and enumerate candidate mechanisms by which postsynaptic neurons may signal retrogradely to determine presynaptic molecular identity and control synaptic dynamics. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Stan A, Pielarski KN, Brigadski T, Wittenmayer N, Fedorchenko O, Gohla A, Lessmann V, Dresbach T, Gottmann K. Essential cooperation of N-cadherin and neuroligin-1 in the transsynaptic control of vesicle accumulation. Proc Natl Acad Sci U S A. 2010;107:11116–11121. doi: 10.1073/pnas.0914233107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Togashi H, Abe K, Mizoguchi A, Takaoka K, Chisaka O, Takeichi M. Cadherin regulates dendritic spine morphogenesis. Neuron. 2002;35:77–89. doi: 10.1016/s0896-6273(02)00748-1. [DOI] [PubMed] [Google Scholar]
- 18.Sara Y, Biederer T, Atasoy D, Chubykin A, Mozhayeva MG, Sudhof TC, Kavalali ET. Selective capability of SynCAM and neuroligin for functional synapse assembly. J Neurosci. 2005;25:260–270. doi: 10.1523/JNEUROSCI.3165-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Hoppa MB, Lana B, Margas W, Dolphin AC, Ryan TA. Alpha2delta expression sets presynaptic calcium channel abundance and release probability. Nature. 2012;486:122–125. doi: 10.1038/nature11033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Kim SH, Ryan TA. Balance of calcineurin Aalpha and CDK5 activities sets release probability at nerve terminals. J Neurosci. 2013;33:8937–8950. doi: 10.1523/JNEUROSCI.4288-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Han Y, Kaeser PS, Sudhof TC, Schneggenburger R. RIM determines Ca(2)+ channel density and vesicle docking at the presynaptic active zone. Neuron. 2011;69:304–316. doi: 10.1016/j.neuron.2010.12.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Hoppa MB, Gouzer G, Armbruster M, Ryan TA. Control and plasticity of the presynaptic action potential waveform at small CNS nerve terminals. Neuron. 2014;84:778–789. doi: 10.1016/j.neuron.2014.09.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23•.Kawaguchi S-y, Sakaba T. Control of inhibitory synaptic outputs by low excitability of axon terminals revealed by direct recording. Neuron. 2015 doi: 10.1016/j.neuron.2015.02.013. The authors of this study directly record from the axon terminals of cerebellar Purkinje neurons and basket cells and demonstrate that STP is strongly influenced by action potential attenuation though to result from lower Na+ and higher K+ channel expression in terminals. Action potential attenuation is less pronounced at basket cell terminals than Purkinje cell terminals, suggesting cell-type and/or synapse-type-specific regulation of axon terminal excitability. [DOI] [PubMed] [Google Scholar]
- 24.de Wit J, Ghosh A. Control of neural circuit formation by leucine-rich repeat proteins. Trends Neurosci. 2014;37:539–550. doi: 10.1016/j.tins.2014.07.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Reyes A, Sakmann B. Developmental switch in the short-term modification of unitary EPSPs evoked in layer 2/3 and layer 5 pyramidal neurons of rat neocortex. J Neurosci. 1999;19:3827–3835. doi: 10.1523/JNEUROSCI.19-10-03827.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Wang L, Fontanini A, Maffei A. Experience-dependent switch in sign and mechanisms for plasticity in layer 4 of primary visual cortex. J Neurosci. 2012;32:10562–10573. doi: 10.1523/JNEUROSCI.0622-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Cheetham CE, Fox K. Presynaptic development at L4 to l2/3 excitatory synapses follows different time courses in visual and somatosensory cortex. J Neurosci. 2010;30:12566–12571. doi: 10.1523/JNEUROSCI.2544-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Angulo MC, Staiger JF, Rossier J, Audinat E. Developmental synaptic changes increase the range of integrative capabilities of an identified excitatory neocortical connection. J Neurosci. 1999;19:1566–1576. doi: 10.1523/JNEUROSCI.19-05-01566.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Crins TT, Rusu SI, Rodriguez-Contreras A, Borst JG. Developmental changes in short-term plasticity at the rat calyx of Held synapse. J Neurosci. 2011;31:11706–11717. doi: 10.1523/JNEUROSCI.1995-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Wasling P, Hanse E, Gustafsson B. Developmental changes in release properties of the CA3-CA1 glutamate synapse in rat hippocampus. J Neurophysiol. 2004;92:2714–2724. doi: 10.1152/jn.00464.2004. [DOI] [PubMed] [Google Scholar]
- 31••.Larsen RS, Smith IT, Miriyala J, Han JE, Corlew RJ, Smith SL, Philpot BD. Synapse-specific control of experience-dependent plasticity by presynaptic NMDA receptors. Neuron. 2014;83:879–893. doi: 10.1016/j.neuron.2014.07.039. This study demonstrates that sensory deprivation uniquely modulates both STP and long-term plasticity at L4 to L2/3 visual cortical synapses. Sensory deprivation influences the expression of GluN3A-containing preNMDARs, resulting in changes in STP and in the ability to induce timing-dependent LTD at these synapses. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32••.Lu J, Tucciarone J, Lin Y, Huang ZJ. Input-specific maturation of synaptic dynamics of parvalbumin interneurons in primary visual cortex. Proc Natl Acad Sci U S A. 2014;111:16895–16900. doi: 10.1073/pnas.1400694111. Using rabies virus monosynaptic tracing, optogenetic stimulation of intracortical inputs, and measurements of STP at long-range and short-range inputs to parvalbumin-positive cortical neurons, the authors map the connections to these neurons and demonstrate a developmental shift in STP that occurs selectively at local connections. This developmental change at local inputs requires changes in the probability of presynaptic release as well as a change in the postsynaptic expression of calcium-permeable AMPARs. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Larsen RS, Corlew RJ, Henson MA, Roberts AC, Mishina M, Watanabe M, Lipton SA, Nakanishi N, Perez-Otano I, Weinberg RJ, et al. NR3A-containing NMDARs promote neurotransmitter release and spike timing-dependent plasticity. Nat Neurosci. 2011;14:338–344. doi: 10.1038/nn.2750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Greig LC, Woodworth MB, Galazo MJ, Padmanabhan H, Macklis JD. Molecular logic of neocortical projection neuron specification, development and diversity. Nat Rev Neurosci. 2013;14:755–769. doi: 10.1038/nrn3586. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Rozov A, Burnashev N. Polyamine-dependent facilitation of postsynaptic AMPA receptors counteracts paired-pulse depression. Nature. 1999;401:594–598. doi: 10.1038/44151. [DOI] [PubMed] [Google Scholar]
- 36.Rozov A, Jerecic J, Sakmann B, Burnashev N. AMPA receptor channels with long-lasting desensitization in bipolar interneurons contribute to synaptic depression in a novel feedback circuit in layer 2/3 of rat neocortex. J Neurosci. 2001;21:8062–8071. doi: 10.1523/JNEUROSCI.21-20-08062.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Banitt Y, Martin KA, Segev I. Depressed responses of facilitatory synapses. J Neurophysiol. 2005;94:865–870. doi: 10.1152/jn.00689.2004. [DOI] [PubMed] [Google Scholar]
- 38.Chen WX, Buonomano DV. Developmental shift of short-term synaptic plasticity in cortical organotypic slices. Neuroscience. 2012;213:38–46. doi: 10.1016/j.neuroscience.2012.04.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Urban-Ciecko J, Wen JA, Parekh PK, Barth AL. Experience-dependent regulation of presynaptic NMDARs enhances neurotransmitter release at neocortical synapses. Learn Mem. 2014;22:47–55. doi: 10.1101/lm.035741.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Bender KJ, Allen CB, Bender VA, Feldman DE. Synaptic basis for whisker deprivation-induced synaptic depression in rat somatosensory cortex. J Neurosci. 2006;26:4155–4165. doi: 10.1523/JNEUROSCI.0175-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Cheetham CE, Fox K. The role of sensory experience in presynaptic development is cortical area specific. J Physiol. 2011;589:5691–5699. doi: 10.1113/jphysiol.2011.218347. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Finnerty GT, Roberts LS, Connors BW. Sensory experience modifies the short-term dynamics of neocortical synapses. Nature. 1999;400:367–371. doi: 10.1038/22553. [DOI] [PubMed] [Google Scholar]
- 43.Takesian AE, Kotak VC, Sanes DH. Presynaptic GABA(B) receptors regulate experience-dependent development of inhibitory short-term plasticity. J Neurosci. 2010;30:2716–2727. doi: 10.1523/JNEUROSCI.3903-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Maffei A, Turrigiano GG. Multiple modes of network homeostasis in visual cortical layer 2/3. J Neurosci. 2008;28:4377–4384. doi: 10.1523/JNEUROSCI.5298-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Sjöström PJ, Turrigiano GG, Nelson SB. Neocortical LTD via coincident activation of presynaptic NMDA and cannabinoid receptors. Neuron. 2003;39:641–654. doi: 10.1016/s0896-6273(03)00476-8. [DOI] [PubMed] [Google Scholar]
- 46.Bidoret C, Ayon A, Barbour B, Casado M. Presynaptic NR2A-containing NMDA receptors implement a high-pass filter synaptic plasticity rule. Proc Natl Acad Sci U S A. 2009;106:14126–14131. doi: 10.1073/pnas.0904284106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Brasier DJ, Feldman DE. Synapse-specific expression of functional presynaptic NMDA receptors in rat somatosensory cortex. J Neurosci. 2008;28:2199–2211. doi: 10.1523/JNEUROSCI.3915-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Corlew R, Wang Y, Ghermazien H, Erisir A, Philpot BD. Developmental switch in the contribution of presynaptic and postsynaptic NMDA receptors to long-term depression. J Neurosci. 2007;27:9835–9845. doi: 10.1523/JNEUROSCI.5494-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Markram H, Gerstner W, Sjöström PJ. A history of spike-timing-dependent plasticity. Front Synaptic Neurosci. 2011;3:4. doi: 10.3389/fnsyn.2011.00004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Nabavi S, Fox R, Proulx CD, Lin JY, Tsien RY, Malinow R. Engineering a memory with LTD and LTP. Nature. 2014;511:348–352. doi: 10.1038/nature13294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Larsen RS, Rao D, Manis PB, Philpot BD. STDP in the developing sensory neocortex. Front Synaptic Neurosci. 2010;2:9. doi: 10.3389/fnsyn.2010.00009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52•.Oh WC, Hill TC, Zito K. Synapse-specific and size-dependent mechanisms of spine structural plasticity accompanying synaptic weakening. Proc Natl Acad Sci U S A. 2013;110:E305–E312. doi: 10.1073/pnas.1214705110. This study shows that, at large mature spines, LTD requires both NMDAR and mGluR activation, while at small immature spines, LTD requires only NMDAR activation. Whether mature inputs are correlated with specific input types is unknown, however. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Holbro N, Grunditz A, Oertner TG. Differential distribution of endoplasmic reticulum controls metabotropic signaling and plasticity at hippocampal synapses. Proc Natl Acad Sci U S A. 2009;106:15055–15060. doi: 10.1073/pnas.0905110106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Kasai H, Fukuda M, Watanabe S, Hayashi-Takagi A, Noguchi J. Structural dynamics of dendritic spines in memory and cognition. Trends Neurosci. 2010;33:121–129. doi: 10.1016/j.tins.2010.01.001. [DOI] [PubMed] [Google Scholar]
- 55.Hulme SR, Jones OD, Abraham WC. Emerging roles of metaplasticity in behaviour and disease. Trends Neurosci. 2013;36:353–362. doi: 10.1016/j.tins.2013.03.007. [DOI] [PubMed] [Google Scholar]
- 56.Paoletti P, Bellone C, Zhou Q. NMDA receptor subunit diversity: impact on receptor properties, synaptic plasticity and disease. Nat Rev Neurosci. 2013;14:383–400. doi: 10.1038/nrn3504. [DOI] [PubMed] [Google Scholar]
- 57.Lee MC, Yasuda R, Ehlers MD. Metaplasticity at single glutamatergic synapses. Neuron. 2010;66:859–870. doi: 10.1016/j.neuron.2010.05.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Colgan LA, Yasuda R. Plasticity of dendritic spines: subcompartmentalization of signaling. Annu Rev Physiol. 2014;76:365–385. doi: 10.1146/annurev-physiol-021113-170400. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Beique JC, Na Y, Kuhl D, Worley PF, Huganir RL. Arc-dependent synapse-specific homeostatic plasticity. Proc Natl Acad Sci U S A. 2011;108:816–821. doi: 10.1073/pnas.1017914108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Sjöström PJ, Häusser M. A cooperative switch determines the sign of synaptic plasticity in distal dendrites of neocortical pyramidal neurons. Neuron. 2006;51:227–238. doi: 10.1016/j.neuron.2006.06.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Markram H, Lübke J, Frotscher M, Roth A, Sakmann B. Physiology and anatomy of synaptic connections between thick tufted pyramidal neurones in the developing rat neocortex. J Physiol. 1997;500(Pt 2):409–440. doi: 10.1113/jphysiol.1997.sp022031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62•.Bartol TM, Bromer C, Kinney JP, Chirillo MA, Bourne JN, Harris KM, Sejnowski TJ. Hippocampal spine head sizes are highly precise. bioRxiv. 2015 This study shows that synaptic weight is correlated for synaptic contacts sharing the same axon and dendrite, showing that synaptic plasticity may be specific to individual synaptic contacts. [Google Scholar]
- 63.Sorra KE, Harris KM. Occurrence and three-dimensional structure of multiple synapses between individual radiatum axons and their target pyramidal cells in hippocampal area CA1. J Neurosci. 1993;13:3736–3748. doi: 10.1523/JNEUROSCI.13-09-03736.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Corlew R, Brasier DJ, Feldman DE, Philpot BD. Presynaptic NMDA receptors: newly appreciated roles in cortical synaptic function and plasticity. Neuroscientist. 2008;14:609–625. doi: 10.1177/1073858408322675. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Park H, Popescu A, Poo MM. Essential role of presynaptic NMDA receptors in activity-dependent BDNF secretion and corticostriatal LTP. Neuron. 2014;84:1009–1022. doi: 10.1016/j.neuron.2014.10.045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Rodriguez-Moreno A, Paulsen O. Spike timing-dependent long-term depression requires presynaptic NMDA receptors. Nat Neurosci. 2008;11:744–745. doi: 10.1038/nn.2125. [DOI] [PubMed] [Google Scholar]
- 67••.Banerjee A, Gonzalez-Rueda A, Sampaio-Baptista C, Paulsen O, Rodriguez-Moreno A. Distinct mechanisms of spike timing-dependent LTD at vertical and horizontal inputs onto L2/3 pyramidal neurons in mouse barrel cortex. Physiol Rep. 2014;2:e00271. doi: 10.1002/phy2.271. The authors find that STDP at ascending layer 4 and intralaminar layer 2/3 inputs onto layer 2/3 pyramidal cells in mouse barrel cortex have different timing requirements and rely on different molecular mechanisms. This is a prime example of two different STSP learning rules existing side-by-side at excitatory synapses onto the same postsynaptic cell. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Froemke RC, Poo MM, Dan Y. Spike-timing-dependent synaptic plasticity depends on dendritic location. Nature. 2005;434:221–225. doi: 10.1038/nature03366. [DOI] [PubMed] [Google Scholar]
- 69.Banerjee A, Meredith RM, Rodriguez-Moreno A, Mierau SB, Auberson YP, Paulsen O. Double dissociation of spike timing-dependent potentiation and depression by subunit-preferring NMDA receptor antagonists in mouse barrel cortex. Cereb Cortex. 2009;19:2959–2969. doi: 10.1093/cercor/bhp067. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Petilla Interneuron Nomenclature G. Ascoli GA, Alonso-Nanclares L, Anderson SA, Barrionuevo G, Benavides-Piccione R, Burkhalter A, Buzsaki G, Cauli B, Defelipe J, et al. Petilla terminology: nomenclature of features of GABAergic interneurons of the cerebral cortex. Nat Rev Neurosci. 2008;9:557–568. doi: 10.1038/nrn2402. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.DeFelipe J, Lopez-Cruz PL, Benavides-Piccione R, Bielza C, Larranaga P, Anderson S, Burkhalter A, Cauli B, Fairen A, Feldmeyer D, et al. New insights into the classification and nomenclature of cortical GABAergic interneurons. Nat Rev Neurosci. 2013;14:202–216. doi: 10.1038/nrn3444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Kepecs A, Fishell G. Interneuron cell types are fit to function. Nature. 2014;505:318–326. doi: 10.1038/nature12983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73••.Chabrol FP, Arenz A, Wiechert MT, Margrie TW, DiGregorio DA. Synaptic diversity enables temporal coding of coincident multisensory inputs in single neurons. Nat Neurosci. 2015;18:718–727. doi: 10.1038/nn.3974. This study shows that cerebellar granule cells receive inputs from different modalities that have distinct activity patterns. Importantly, short-term STSP enables individual granule cells to temporally code multimodal input, which facilitates pattern separation. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Seshadri S, Zeledon M, Sawa A. Synapse-specific contributions in the cortical pathology of schizophrenia. Neurobiol Dis. 2013;53:26–35. doi: 10.1016/j.nbd.2013.01.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Wilke SA, Raam T, Antonios JK, Bushong EA, Koo EH, Ellisman MH, Ghosh A. Specific disruption of hippocampal mossy fiber synapses in a mouse model of familial Alzheimer’s disease. PLOS ONE. 2014;9:e84349. doi: 10.1371/journal.pone.0084349. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Patel AB, Loerwald KW, Huber KM, Gibson JR. Postsynaptic FMRP promotes the pruning of cell-to-cell connections among pyramidal neurons in the L5A neocortical network. J Neurosci. 2014;34:3413–3418. doi: 10.1523/JNEUROSCI.2921-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Wallace ML, Burette AC, Weinberg RJ, Philpot BD. Maternal loss of Ube3a produces an excitatory/inhibitory imbalance through neuron type-specific synaptic defects. Neuron. 2012;74:793–800. doi: 10.1016/j.neuron.2012.03.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Sun HY, Dobrunz LE. Presynaptic kainate receptor activation is a novel mechanism for target cell-specific short-term facilitation at Schaffer collateral synapses. J Neurosci. 2006;26:10796–10807. doi: 10.1523/JNEUROSCI.2746-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Eliasmith C, Trujillo O. The use and abuse of large-scale brain models. Curr Opin Neurobiol. 2014;25:1–6. doi: 10.1016/j.conb.2013.09.009. [DOI] [PubMed] [Google Scholar]
- 80.Choudhury S, Fishman JR, McGowan ML, Juengst ET. Big data, open science and the brain: lessons learned from genomics. Front Hum Neurosci. 2014;8:239. doi: 10.3389/fnhum.2014.00239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Lüscher C, Isaac JT. The synapse: center stage for many brain diseases. J Physiol. 2009;587:727–729. doi: 10.1113/jphysiol.2008.167742. [DOI] [PMC free article] [PubMed] [Google Scholar]



