Skip to main content
Howard Hughes Medical Institute Author Manuscripts logoLink to Howard Hughes Medical Institute Author Manuscripts
. Author manuscript; available in PMC: 2010 Jul 8.
Published in final edited form as: Curr Opin Neurobiol. 2000 Apr;10(2):172–179. doi: 10.1016/s0959-4388(00)00076-3

Natural patterns of activity and long-term synaptic plasticity

Ole Paulsen *, Terrence J Sejnowski
PMCID: PMC2900254  NIHMSID: NIHMS217351  PMID: 10753798

Abstract

Long-term potentiation (LTP) of synaptic transmission is traditionally elicited by massively synchronous, high-frequency inputs, which rarely occur naturally. Recent in vitro experiments have revealed that both LTP and long-term depression (LTD) can arise by appropriately pairing weak synaptic inputs with action potentials in the postsynaptic cell. This discovery has generated new insights into the conditions under which synaptic modification may occur in pyramidal neurons in vivo. First, it has been shown that the temporal order of the synaptic input and the postsynaptic spike within a narrow temporal window determines whether LTP or LTD is elicited, according to a temporally asymmetric Hebbian learning rule. Second, backpropagating action potentials are able to serve as a global signal for synaptic plasticity in a neuron compared with local associative interactions between synaptic inputs on dendrites. Third, a specific temporal pattern of activity — postsynaptic bursting — accompanies synaptic potentiation in adults.

Introduction

Acts of recollection, as they occur in experience, are due to the fact that one movement has by nature another that succeeds it in regular order. If this order be necessary, whenever a subject experiences the former of two movements thus connected, it will (invariably) experience the latter.

περὶ μνήμηϛ καὶ ἀναμνησέωϛ

Aristotle, Fourth century B.C.

De memoria et reminiscentia

[Title translation: On memory and reminiscences]

This citation from Aristotle highlights two properties of long-term memories for facts and events: their associative nature and their temporal order. These properties were incorporated into an influential proposal for synaptic plasticity made by Donald Hebb in 1949. He suggested that reverberatory activity in transient assemblies of neurons carries a memory trace that becomes permanently laid down as changes in synaptic weights when a presynaptic neuron repeatedly or persistently takes part in firing the postsynaptic cell [1]. Major research efforts that are underway to explain long-term memory have uncovered mechanisms that are consistent with both the associative nature and the temporal order emphasized in Hebb's proposal.

The first biological mechanism discovered that could potentially support Hebb's learning rule was long-term potentiation (LTP) of synaptic transmission, as described by Bliss and Lømo [2,3]. The original protocol for inducing LTP was high-frequency stimulation of presynaptic neurons. However, the highly synchronous population activity required to induce this type of LTP has never been observed during learning in vivo. An important question is whether synaptic potentiation could be induced by more natural activity patterns based on the relative timing of presynaptic and postsynaptic activity as originally suggested by Hebb.

In 1994, Stuart and Sakmann [4] reported that action potentials can backpropagate in the dendrites of cortical pyramidal neurons. More recently, it has been shown that backpropagating spikes could directly serve as an associative signal for LTP induction under some experimental conditions [5,6]. These papers have raised interest in the induction criteria for LTP, using behaviorally relevant stimuli. This paper briefly reviews the type of neuronal activity that can be recorded during learning episodes in vivo, and, with this background, discusses three fundamental questions about synaptic learning rules based on recent experimental evidence obtained in vitro:

  1. Are synaptic learning rules temporal coincidence rules, or is the temporal order of presynaptic and postsynaptic activity important?

  2. Is the postsynaptic induction of long-term plasticity controlled by an associative signal localized in dendritic segments, or is there a signal that is global to the neuron?

  3. Is all successful information transfer through a neuron associated with updating of synaptic weights, or is there a specific type of activity that occurs during synaptic modification?

In this review, we focus on the postsynaptic neuron in inducing and regulating synaptic plasticity and emphasize the predictive, in addition to the associative, nature of Hebbian learning.

Neuronal activity during behavioral learning

The conditions for synaptic plasticity in the behaving animal must be sought among the activity patterns that occur during learning. The hippocampus is a structure of critical importance in memory for facts and events [7]. In addition to its established role in the processing of spatial information in rodents [8], recent results indicate that hippocampal neurons can also encode non-spatial information [9,10] and take part in several memory processes [11]. Recent experiments have shed light on the nature of the encoding system in the hippocampus. Although memory retrieval can be impaired by small localized lesions within the hippocampus, the induction of new memories is not [12••]. Also, repeated tetanization in a limited amount of hippocampal tissue can disrupt learning that would otherwise be supported [13]. These findings raise the question of what the natural conditions are that lead to synaptic modifications that could support such a distributed encoding of memories.

Refinement of multielectrode recording techniques (multiple tetrodes and multisite silicon probes) has given us new insights into the neuronal activity that occurs during learning-related behavior in rats. Exploratory learning is associated with characteristic slow rhythmic activity in extracellularly recorded potentials [14]. Interestingly, similar oscillations in the electroencephalogram (EEG) are seen during memory tasks in humans [15]. This so-called theta activity is coexistent with faster oscillations at gamma frequencies (30–100 Hz) [16]. In hippocampal slice experiments, theta-like oscillations can be induced by carbachol, a cholinergic agonist, and 40 Hz activity can co-occur over a range of carbachol concentrations [17,18]. The occurrence of rhythmic oscillatory activity may engage learning mechanisms that depend on relative timing of activity in the presynaptic and postsynaptic neurons, since the rhythm will naturally organize the timing of the key elements of the neural circuit.

Hippocampal pyramidal neurons fire, on average, relatively infrequently (< 1 Hz) but their spikes are phase-related to the external theta field oscillation. They may be silent, fire single spikes or brief high-frequency trains of spikes (‘bursts’) during each cycle [19]. Pyramidal neurons fire bursts primarily when the animal is in a specific location in space (the ‘place field’ of that particular neuron) [20,21]. Pyramidal layer interneurons have a higher firing rate, and fire often on every cycle during a theta rhythm [19]. In behaving animals, interneurons fire at the same phase of theta as do pyramidal neurons [19], in contrast to what has been reported in anaesthetized animals, where basket interneurons and pyramidal cells fire at opposite phases of the external theta oscillations [22]. As spikes in both the pyramidal neurons and the interneurons are phase-related to gamma oscillation [16,23], and as interneurons fire on the same phase of theta as pyramidal cells, they might control the exact timing of pyramidal cell firing in 40 Hz subcycles of each theta cycle [24]. There is now evidence that action potentials can backpropagate in pyramidal neuron dendrites in vivo [25], and backpropagating action potentials can, under some circumstances, trigger Ca2+-spike-associated bursts [26].

These observations raise the possibility that burst firing and phase-related firing patterns could support the induction of synaptic potentiation. Recent in vitro experiments have addressed this issue.

Temporal constraint on pre- and postsynaptic activity: an asymmetric Hebbian learning rule

The Hebbian learning rule has often been interpreted to mean that synaptic potentiation should occur as a consequence of coincident activity in presynaptic and postsynaptic neurons. However, Hebb's original suggestion incorporated a temporal constraint, namely that presynaptic activity must precede the activity in the postsynaptic element for potentiation to occur [1].

Experiments setting out to test the effects of relative timing of spikes directly using single backpropagating action potentials as the postsynaptic associative signal have uncovered that the precise temporal order of the pre- and postsynaptic signals of pairing is important [5,27,28••,29]. In developing neurons, the postsynaptic action potential needs to take place within a narrow time window (within ~20 ms) following the presynaptic action potential for robust potentiation to occur (as reported in studies in hippocampal slice cultures [27], cultured hippocampal pyramidal neurons [28••] and developing frog tectal neurons in vivo [29]). Moreover, when the pairing was reversed, so that the postsynaptic action potential occurred immediately before the presynaptic action potential, depression was induced [5,27,28••,29]. Thus, at least in developing systems, a correlational learning rule exists at the synaptic level where the sign of synaptic change depends on the relative timing of the presynaptic input and the postsynaptic action potential. These rules have been studied in networks of cultured neurons, where poly-synaptic pathways are modified according to this temporally asymmetric Hebbian rule [30].

In computer simulations of recurrent hippocampal networks, temporally asymmetric Hebbian synaptic plasticity supports sequence learning [31,32]. This occurs because the learning rule tends to wire together neurons that form causal chains; that is, if neuron A fires before neuron B and they represent two sequential sensory or motor states, the connection between them will strengthen so that in the future, neuron B is more likely to fire after neuron A. A sequence of patterns can be stored in an oscillating network model with recurrent excitatory synapses, as occurs in the CA3 region of the hippocampus where 5–8 Hz theta oscillations are observed in vivo [24]. The sequence could then be rapidly retrieved from the network in 40 Hz subcycles of the slower theta oscillation [18,33].

The motion of visual stimuli provides the visual cortex with a sequence of highly correlated inputs. The development of direction selectivity in developing visual cortical neurons can be modeled with an asymmetric Hebbian rule implemented in a recurrent network [34••].

Backpropagating action potentials as a global neuronal associative signal

The induction of LTP at some excitatory synapses on pyramidal neurons depends on the activation of NMDA receptors [3]. NMDA receptors are thought to serve as molecular coincidence detectors, requiring the presynaptic release of glutamate immediately followed by postsynaptic depolarization. Where does the postsynaptic depolarization originate? The traditional view is that a ‘strong’ synaptic input can depolarize the local dendritic branch sufficiently to enable the activation of NMDA receptors at ‘weak’ inputs on neighboring synapses [3]. However, back-propagating action potentials could also provide the postsynaptic associative signal [35]. A backpropagating action potential would serve as a global dendritic associative signal, reaching a large fraction — and potentially all — of the synapses on a single neuron. The extent of the influence of the backpropagating action potential could be subject to control by ion channels and synaptic inhibition. The physiological control of dendritic backpropagation of action potentials has recently been reviewed elsewhere [36]. Figure 1 illustrates the difference between the concepts of local cooperativity in the dendrites and global backpropagating signals.

Figure 1.

Figure 1

Two models for induction of synaptic modifications in the hippocampus. (a) Conventional model for induction of LTP. A ‘strong’ input (large number of synchronously active afferent input fibers) produces a local dendritic depolarization that unblocks NMDA receptors. Synaptically released glutamate in neighboring excitatory synapses, concurrently active with the strong input, can activate the NMDA receptor providing the necessary Ca2+ signal for induction of LTP. (b) New model for induction of synaptic modifications based on backpropagating action potentials. In this scenario, postsynaptic action potentials provide a global signal in the neuron, allowing all synapses onto this neuron to be modified, according to their exact timing relative to the postsynaptic action potentials (temporally asymmetric Hebbian learning rule). Postsynaptic bursting signals potentiation in recently active synapses.

Induction of LTP depends on a postsynaptic Ca2+ increase [3]. It has been widely assumed that the permeation of Ca2+ through NMDA receptors provides the Ca2+ signal necessary for the induction of LTP [3]. However, voltage-gated Ca2+ channels and Ca2+-induced Ca2+ release from intracellular stores may also contribute to the Ca2+ transient observed in dendritic spines following different types of activation [3740]. Confocal imaging and two- and multi-photon microscopy have been used to measure Ca2+ transients in single dendritic spines during synaptic activation and following backpropagation of action potentials [3740]. Single backpropagating action potentials produce a small Ca2+ influx [41], while the Ca2+ transient following pairing of synaptic activation and backpropagating action potentials depends critically on the relative timing of the backpropagating action potentials and the synaptic activation. The spinous Ca2+ transient evoked by pairing an excitatory postsynaptic potential (EPSP) and a backpropagating action potential was larger if the action potential followed the EPSP than if it preceded it [37].

These observations can be modeled using MCell, a computer program which simulates the diffusion of neurotransmitter molecules and subcellular signaling [42,43]. The rise and fall of the backpropagating action potential sets the time scale for the rapid depolarization pulse that occurs at distal synapses. This pulse, in turn, determines the conditions for massive entry of Ca2+ into the spine. The time course for glutamate binding to the NMDA receptor is much longer than the width of the backpropagating action potential, so the dynamics of Ca2+, and the binding of Ca2+ to Ca2+-binding proteins, depend in a highly nonlinear way on the relative timing of these two signals [43]. As a consequence of these biophysical mechanisms, the change in the synaptic strength depends not on the coincidence of presynaptic with postsynaptic activity, but, to a first approximation, with the time derivative of the postsynaptic activity [34••]. This temporally asymmetric form of the Hebbian learning rule has profound consequences for learning in highly recurrent networks of neurons, such as those found in the neocortex and area CA3 of the hippocampus.

The nature of the postsynaptic associative signal in adult animals

In cultured hippocampal neurons, repeated pairings of single pre- and postsynaptic action potentials are sufficient to induce synaptic potentiation [27,28••]. This would be expected to occur under normal developmental conditions during establishment of the network architecture, but does it also apply to hippocampal neurons in adult animals? Apparently not, according to two recent studies in hippocampal slices prepared acutely from adult mice and rats [44••,45••] where pairings of synaptic inputs with single postsynaptic action potentials was not sufficient to induce LTP; however, pairing with postsynaptic bursting activity did induce LTP.

A standard procedure for inducing synaptic plasticity in hippocampal slices from adult mice is a long train of relatively low-frequency afferent activity (1–10 Hz). Under these conditions, synaptic potentiation or depression could be induced depending on the frequency of activation, with depression occurring at the lowest frequencies and potentiation above a frequency threshold [46]. Certain genetic modifications can change the frequency threshold between LTP and LTD (long-term depression), and these usually lead to a deficit in performance in memory tasks [46,47,48]. Consistent with theoretical models of developing cortex, which predicted a sliding modification threshold [49], the frequency threshold can be dynamically altered by priming with a high-frequency input [50].

A recent discovery has raised another interesting possibility regarding the nature of the threshold separating LTP from LTD. By monitoring the postsynaptic cell with intracellular recording, it was shown that at 5 Hz afferent activity, induction of LTP depended on postsynaptic burst firing [44••]. A direct comparison of the efficiency of single spikes and bursts in inducing LTP using behaviorally plausible stimuli revealed that bursting was indeed more efficient than single spikes in inducing LTP, and might be necessary for potentiation to occur in slices from adult rats [45••]. In these experiments, postsynaptic bursting was both necessary and sufficient for synaptic potentiation when paired with presynaptic single spikes, whereas presynaptic bursting was neither necessary nor sufficient [45••]. It is conceivable that in the adult bursting satisfies the postsynaptic requirements for NMDA receptor activation, whereas single action potentials do not satisfy the requirements. Thus, a single burst of action potentials may replace the need in developing neurons for repeated pairing of single spikes (Figure 1).

What controls burst firing in neurons? A specific pattern of activity may provide the signal for bursting to occur. It was recently found in neocortical neurons that distal EPSPs that immediately follow backpropagating action potentials (3–7 ms later) can trigger Ca2+ spikes in the dendrites [51••]. Such dendritic Ca2+ spikes are accompanied by burst firing in the soma [51••,52,53]. Direct recording from dendrites of CA1 pyramidal cells in the hippocampus in vivo has confirmed that dendritic Ca2+ spikes are always preceded by fast action potentials [26]. Interestingly, the magnitude of action-potential-induced dendritic Ca2+ transients correlate positively to LTP induction by theta burst stimuli in developing hippocampal neurons [54].

Bursting in hippocampal neurons occurs naturally under two behavioral conditions. First, during active exploration of novel environments, hippocampal neurons that code for the current location in space of the animal typically show bursting activity repeated at an interburst frequency between 5 and 12 Hz [20,21]. These are the hippocampal neurons most likely to be involved in associative learning. Second, during slow-wave sleep, spike sequences detected during behavior are ‘replayed’ on a faster timescale during sharp-wave bursts, and it has been suggested that this activity is associated with the consolidation of memory [55,56••]. It has been reported that LTP can be induced during sharp waves by pairing sharp-wave activity with strong postsynaptic depolarization of individual pyramidal cells [57]. The occurrence of bursting under conditions presumed to be associated with memory induction in vivo along with the requirement of bursting for synaptic potentiation in vitro suggests that in the adult, bursting is a neuronal activity specifically signaling memory-related synaptic potentiation.

A specific postsynaptic activity signaling synaptic potentiation would be interesting for at least two reasons. First, it re-emphasizes the critical role of activity in the postsynaptic neuron for synaptic plasticity to occur. Second, it suggests that at least three logic levels of signaling exist in memory encoding: silence, single spikes transferring information, and bursts signifying changes in synaptic weights. Whereas in developing neurons single spikes provide the adequate signal for laying down the architecture of a network, in the adult a reinforcement signal — bursting triggered by a specific afferent input — might be required for synaptic plasticity to occur.

The existence of bursting as a global signal might have far-reaching implications. Burst firing in the cell body could play a role in synaptic plasticity also by action-potential-dependent regulation of gene expression [58]. But how could a signal from the nucleus reach and modify only the relevant synapses? Recently, Frey and Morris [59] showed that weak tetanization which would normally lead to short-lasting LTP (‘early LTP’) could induce longer-lasting LTP (‘late LTP’) if a separate pathway had been strongly tetanized earlier. Since late LTP requires both protein and mRNA synthesis, induction of LTP may be associated with the setting of a ‘synaptic tag’ at activated synapses, which is capable of sequestering a plasticity-maintenance chemical message from the nucleus [60••]. More recently it has been suggested that the late-LTP signal does not need to be synaptic, but can be replaced by postsynaptic bursting activity [61]. If this is indeed the case, it would be tempting to speculate that the burst firing that is so prevalent during sleep might have a function related to the maintenance of synaptic weights in recently altered synapses (‘consolidation’) and for the structural reorganization of the neuropil, perhaps involving new spines and dendritic branches, which may require gene regulation [6264].

Computational consequences

The demonstration that patterns of activity that occur in vivo during learning can elicit long-term changes of synaptic strengths in vitro makes it more likely that we are getting closer to understanding the mechanisms underlying learning and memory. If so, then some of the conditions accompanying these changes may be important clues to the cellular substrates of the behavioral changes that accompany learning. In particular, the time scale and the temporal asymmetry in the learning rule have important implications for the organization of cortical circuits. The importance of temporal order on a millisecond time scale for eliciting LTP and LTD suggests that precise cellular and molecular mechanisms may regulate spike timing in cortical circuits [65].

The timing of a spike in the postsynaptic neuron divides the excitatory input activities that occur during the +/–20 ms temporal window into two groups: those synapses that contributed to depolarization preceding the spike, and those that cannot have contributed because they were activated after the spike. This causal structure is directly translated into synaptic plasticity by the temporally asymmetric learning rule, which potentiates those synapses that are active immediately preceding the postsynaptic spike and depresses those synapses that are active directly after the spike [5]. This takes into account the temporal order of continuous events in the world, as noted by Aristotle.

The temporally asymmetric Hebbian learning rule implements the temporal difference (TD) learning algorithm in reinforcement learning [31]. TD learning allows predictions to be made about future events in the world as a consequence of experiencing sequences of inputs. For example, in a model system for honeybee foraging based on classical conditioning [66], the synaptic input learns to predict future reward: if the actual reward is greater than predicted, the postsynaptic neuron is depolarized and the synapse strengthens, but if the reward is less than predicted, the postsynaptic neuron is hyperpolarized and the synapse decreases in strength [66]. At the cellular level in the adult animal, ‘reward’ might be signaled by postsynaptic bursting activity, mediated by the ability of distal inputs to trigger dendritic Ca2+ spikes if timed immediately following a backpropagating fast action potential [51••]. Temporal order is also important for classical conditioning, though the time window is a few seconds — two orders of magnitude longer than that found in the cortex and hippocampus. Thus, TD learning can be used by the cortex to memorize long sequences of input states, which might be useful for storing a musical composition, in addition to the much slower learning of strategies for survival in an uncertain environment [67].

Perhaps the most exciting theoretical implication of the temporally asymmetric Hebbian learning rule is its ability to create and stabilize activity patterns in neural assemblies. First, the combination of LTP and LTD achieves a balance which overcomes the problem of synaptic saturation found to occur with learning rules that can only increase the strengths of synapses [68]. Second, rather than requiring a consensus of coupled neurons to fire together in order to wire together, the temporal asymmetry in the learning means that the first neuron to respond to a new pattern could help to recruit other neurons that are slower to respond through temporally asymmetric Hebbian plasticity [69]. Finally, the temporally asymmetric Hebbian learning rule will tend to make persistent patterns of activity in a recurrent neural assembly more persistent, as Hebb first suggested, but ironically this model requires the anti-Hebbian version, in which a synapse decreases in strength when the EPSP precedes the postsynaptic spike, and increases in strength when the EPSP occurs after the spike [70,71]. From discussions at a recent workshop where many of the experimental and theoretical researchers studying time-dependent synaptic plasticity gathered to compare notes on work in progress, it was clear that exciting new developments lie ahead [72].

Conclusions

All of the results reported here were from studies on pyramidal neurons. We do not have comparable knowledge of interneurons, which are highly diverse and may have a variety of different roles in regulating synaptic plasticity. Without an account of mechanisms for plasticity in interneurons it will not be possible to understand how a network of neurons learns new patterns of activity. Another important area that this review has not focused on is presynaptic mechanisms that might also be involved in the maintenance of synaptic plasticity.

The discovery of backpropagating action potentials has refocused attention on the role of the postsynaptic neuron in synaptic plasticity. Evidence is mounting that the relative timing of presynaptic activity and backpropagating action potentials in the postsynaptic cell can induce long-term synaptic changes in hippocampal pyramidal neurons by activity patterns known to occur during learning in vivo. The shift of emphasis to the temporal domain has opened up an exciting new chapter in the theoretical analysis of synaptic plasticity and neural networks, in which the predictive rather than the associative nature of Hebbian learning is being explored [73].

Acknowledgements

We are indebted to Aristotle for sharing his insights into memory mechanisms. We thank Kevin Franks, Paul Ganter, Narayanan Kasthuri, Edvard Moser and Jeremy Seamans for comments on a drafts of the manuscript. The authors are supported by The Wellcome Trust (O Paulsen), The Royal Society (O Paulsen), the Howard Hughes Medical Institute (TJ Sejnowski), and O Paulsen also holds a Junior Research Fellowship in Biological Sciences at Wadham College, Oxford, UK.

Abbreviations

EPSP

excitatory postsynaptic potential

LTD

long-term depression

LTP

long-term potentiation

NMDA

N-methyl-d-aspartate

TD

temporal difference

References and recommended reading

Papers of particular interest, published within the annual period of review, have been highlighted as:

• of special interest

•• of outstanding interest

  • 1.Hebb DO. The Organization of Behavior. John Wiley & Sons; New York: 1949. [Google Scholar]
  • 2.Bliss TVP, Lømo T. Long-lasting potentiation of synaptic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. J Physiol. 1973;232:331–356. doi: 10.1113/jphysiol.1973.sp010273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Bliss TV, Collingridge GL. A synaptic model of memory: long-term potentiation in the hippocampus. Nature. 1993;361:31–39. doi: 10.1038/361031a0. [DOI] [PubMed] [Google Scholar]
  • 4.Stuart GJ, Sakmann B. Active propagation of somatic action potentials into neocortical pyramidal cell dendrites. Nature. 1994;367:69–72. doi: 10.1038/367069a0. [DOI] [PubMed] [Google Scholar]
  • 5.Markram H, Lubke J, Frotscher M, Sakmann B. Regulation of synaptic efficacy by coincidence of postsynaptic APs and EPSPs. Science. 1997;275:213–215. doi: 10.1126/science.275.5297.213. [DOI] [PubMed] [Google Scholar]
  • 6.Magee JC, Johnston D. A synaptically controlled, associative signal for Hebbian plasticity in hippocampal neurons. Science. 1997;275:209–213. doi: 10.1126/science.275.5297.209. [DOI] [PubMed] [Google Scholar]
  • 7.Squire LR. Memory and the hippocampus: a synthesis from findings with rats, monkeys, and humans. Psychol Rev. 1992;99:195–231. doi: 10.1037/0033-295x.99.2.195. [DOI] [PubMed] [Google Scholar]
  • 8.O'Keefe JA, Nadel L. The Hippocampus as a Cognitive Map. Clarendon; Oxford: 1978. [Google Scholar]
  • 9•.Wood ER, Dudchenko PA, Eichenbaum H. The global record of memory in hippocampal neuronal activity. Nature. 1999;397:613–616. doi: 10.1038/17605. [The authors provide compelling evidence for specific hippocampal neuronal activity that is unrelated to particular spatial locations, indicating that the hippocampus has a broader role in memory processing than in spatial navigation] [DOI] [PubMed] [Google Scholar]
  • 10•.Hampson RE, Simeral JD, Deadwyler SA. Distribution of spatial and nonspatial information in dorsal hippocampus. Nature. 1999;402:610–614. doi: 10.1038/45154. [The authors describe a multielectrode recording experiment showing anatomical segregation of spatial information with interleaved representation of nonspatial information in the rat hippocampus] [DOI] [PubMed] [Google Scholar]
  • 11•.Riedel G, Micheau J, Lam AGM, Roloff EVL, Martin SJ, Bridge H, de Hoz L, Poeschel B, McCulloch J, Morris RGM. Reversible neural inactivation reveals hippocampal participation in several memory processes. Nat Neurosci. 1999;2:898–905. doi: 10.1038/13202. [By reversible inactivation of synaptic transmission by intrahippocampal infusion of an antagonist at AMPA/kainate-receptors, the authors show that hippocampal neuronal activity is necessary for both encoding and retrieval of spatial memory in the rat] [DOI] [PubMed] [Google Scholar]
  • 12••.Moser MB, Moser EI. Distributed encoding and retrieval of spatial memory in the hippocampus. J Neurosci. 1998;18:7535–7542. doi: 10.1523/JNEUROSCI.18-18-07535.1998. [To determine whether memory is processed in a localized or distributed manner by the hippocampus, the authors made a number of small discrete lesions in the hippocampus in a rat pretrained in a spatial memory task. These lesions disrupted the retrieval of pre-established memories while sparing the encoding of new memories, suggesting that spatial memory is normally both encoded and retrieved by a widely distributed hippocampal network] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13•.Moser EI, Krobert KA, Moser MB, Morris RGM. Impaired spatial learning after saturation of long-term potentiation. Science. 1998;281:2038–2042. doi: 10.1126/science.281.5385.2038. [Tetanization-induced saturation of LTP correlates with impairment of spatial learning in a Morris water maze task. This result strengthens results of previous studies using antagonists at NMDA receptors to suggest that LTP is necessary for certain aspects of spatial learning in a Morris water maze task] [DOI] [PubMed] [Google Scholar]
  • 14.Buzsáki G, Leung L, Vanderwolf CH. Cellular bases of hippocampal EEG in the behaving rat. Brain Res Rev. 1983;6:139–171. doi: 10.1016/0165-0173(83)90037-1. [DOI] [PubMed] [Google Scholar]
  • 15•.Kahana MJ, Sekuler R, Caplan JB, Kirschen M, Madsen JR. Human theta oscillations exhibit task dependence during virtual maze navigation. Nature. 1999;399:781–784. doi: 10.1038/21645. [Subdural recordings were made from epileptic patients learning to navigate computer-generated mazes. Theta activity similar to that seen during spatial learning and navigation in rodents was seen during this task, suggesting that theta oscillations may have a role in spatial navigation also in humans] [DOI] [PubMed] [Google Scholar]
  • 16•.Penttonen M, Kamondi A, Acsady L, Buzsaki G. Gamma frequency oscillation in the hippocampus of the rat: intracellular analysis in vivo. Eur J Neurosci. 1998;10:718–728. doi: 10.1046/j.1460-9568.1998.00096.x. [The authors describe intracellular recordings made from morphologically identified neurons during gamma oscillations in anesthetized rats] [DOI] [PubMed] [Google Scholar]
  • 17.Fellous J-M, Sejnowski TJ. Cholinergic induction of spontaneous oscillations in the hippocampal slice in the slow (.5–2 Hz), theta (5–12 Hz) and gamma (35–70 Hz) bands. Hippocampus. 2000;10 doi: 10.1002/(SICI)1098-1063(2000)10:2<187::AID-HIPO8>3.0.CO;2-M. in press. [DOI] [PubMed] [Google Scholar]
  • 18.Fisahn A, Pike FG, Buhl EH, Paulsen O. Cholinergic induction of network oscillations at 40 Hz in the hippocampus in vitro. Nature. 1998;394:186–189. doi: 10.1038/28179. [DOI] [PubMed] [Google Scholar]
  • 19•.Csicsvari J, Hirase H, Czurko A, Mamiya A, Buzsaki G. Oscillatory coupling of hippocampal pyramidal cells and interneurons in the behaving rat. J Neurosci. 1999;19:274–287. doi: 10.1523/JNEUROSCI.19-01-00274.1999. [The authors studied firing relations of pyramidal neurons and interneurons to theta oscillations and sharp waves recorded in behaving rats. They found that the average firing rate was similar for both types of neurons during theta and non-theta states. Pyramidal neurons were silent, or fired one or more action potentials during each theta cycle, and the participation during sharp-wave bursts was variable. Interneurons fired 20—60° before the population discharge of pyramidal cells during theta oscillations, whereas the activity of interneurons lagged behind the maximum discharge probability of pyramidal neurons by 90° corresponding to 1–2 ms during sharp-wave-associated field ripples] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Otto T, Eichenbaum H, Wiener SI, Wible CG. Learning-related patterns of CA1 spike trains parallel stimulation parameters optimal for inducing hippocampal long-term potentiation. Hippocampus. 1991;1:181–192. doi: 10.1002/hipo.450010206. [DOI] [PubMed] [Google Scholar]
  • 21.O'Keefe J, Recce ML. Phase relationship between hippocampal place units and the EEG theta rhythm. Hippocampus. 1993;3:317–330. doi: 10.1002/hipo.450030307. [DOI] [PubMed] [Google Scholar]
  • 22.Ylinen A, Soltesz I, Bragin A, Penttonen M, Sik A, Buzsáki G. Intracellular correlates of hippocampal theta rhythm in identified pyramidal cells, granule cells, and basket cells. Hippocampus. 1995;5:78–90. doi: 10.1002/hipo.450050110. [DOI] [PubMed] [Google Scholar]
  • 23.Tsodyks MV, Skaggs WE, Sejnowski TJ, McNaughton BL. Paradoxical effects of external modulation of inhibitory interneurons. J Neurosci. 1997;17:4382–4388. doi: 10.1523/JNEUROSCI.17-11-04382.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24•.Lisman JE. Relating hippocampal circuitry to function: recall of memory sequences by reciprocal dentate-CA3 interactions. Neuron. 1999;22:233–242. doi: 10.1016/s0896-6273(00)81085-5. [The author describes a new model for storage and retrieval of memory sequences in the hippocampus by reciprocal connections between two recurrent networks, the dentate and the CA3. This model represents a step towards trying to link cellular, network and behavioural phenomena] [DOI] [PubMed] [Google Scholar]
  • 25.Buzsaki G, Penttonen M, Nadasdy Z, Bragin A. Pattern and inhibition-dependent invasion of pyramidal cell dendrites by fast spikes in the hippocampus in vivo. Proc Natl Acad Sci USA. 1996;93:9921–9925. doi: 10.1073/pnas.93.18.9921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26•.Kamondi A, Acsady L, Buzsaki G. Dendritic spikes are enhanced by cooperative network activity in the intact hippocampus. J Neurosci. 1998;18:3919–3928. doi: 10.1523/JNEUROSCI.18-10-03919.1998. [Dendritic action potentials were studied by intracellular sharp electrode recordings from morphologically identified sites in the apical dendrites of CA1 pyramidal neurons in anaesthetised rats. The authors found that dendritic fast action potentials decreased in amplitude as a function of distance from the soma, and that dendritic Ca2+ spikes were always preceded by large-amplitude fast spikes] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27•.Debanne D, Gahwiler BH, Thompson SM. Long-term synaptic plasticity between pairs of individual CA3 pyramidal cells in rat hippocampal slice cultures. J Physiol. 1998;507:237–247. doi: 10.1111/j.1469-7793.1998.237bu.x. [The authors describe LTP and LTD of excitatory connections between pairs of CA3 pyramidal cells induced by single pre- and postsynaptic action potentials in hippocampal slice cultures] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28••.Bi GQ, Poo MM. Synaptic modifications in cultured hippocampal neurons: dependence on spike timing, synaptic strength, and postsynaptic cell type. J Neurosci. 1998;18:10464–10472. doi: 10.1523/JNEUROSCI.18-24-10464.1998. [This is a compelling demonstration of an asymmetric Hebbian rule for NMDA-receptor-dependent synaptic modification of excitatory synapses between cultured glutamatergic (but not GABAergic) hippocampal neurons. Repeated pairing of postsynaptic spiking within 20 ms after presynaptic activation resulted in LTP, whereas postsynaptic spiking within 20 ms before presynaptic activation led to LTD] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29•.Zhang LI, Tao HW, Holt CE, Harris WA, Poo MM. A critical window for cooperation and competition among developing retinotectal synapses. Nature. 1998;395:37–44. doi: 10.1038/25665. [The authors demonstrate that excitatory synapses in the developing frog visual system can undergo activity-dependent potentiation or depression depending on the temporal order of activation of pre- and postsynaptic neurons] [DOI] [PubMed] [Google Scholar]
  • 30•.Bi G, Poo MM. Distributed synaptic modification in neural networks induced by patterned stimulation. Nature. 1999;401:792–796. doi: 10.1038/44573. [The authors performed experiments confirming the rule demonstrated in [28••] for polysynaptic pathways in cultured hippocampal neurons] [DOI] [PubMed] [Google Scholar]
  • 31.Abbott LF, Blum KI. Functional significance of long-term potentiation for sequence learning and prediction. Cereb Cortex. 1996;6:406–416. doi: 10.1093/cercor/6.3.406. [DOI] [PubMed] [Google Scholar]
  • 32.Blum KI, Abbott LF. A model of spatial map formation in the hippocampus of the rat. Neural Comp. 1996;8:85–93. doi: 10.1162/neco.1996.8.1.85. [DOI] [PubMed] [Google Scholar]
  • 33.Lisman J. What makes the brain's tickers tock. Nature. 1998;394:132–133. doi: 10.1038/28061. [DOI] [PubMed] [Google Scholar]
  • 34••.Rao RPN, Sejnowski TJ. Predictive learning of temporal sequences in recurrent neocortical circuits. In: Solla SA, Leen TK, Müller K-R, editors. Advances in Neural Information Processing Systems 12. MIT Press; Cambridge, MA: 2000. in press. [A network with recurrently connected neurons displaying asymmetric Hebbian plasticity can learn sequences and, in particular, can develop direction selectivity similar to that observed in complex cells in visual cortex] [Google Scholar]
  • 35•.Linden DJ. The return of the spike: postsynaptic action potentials and the induction of LTP and LTD. Neuron. 1999;22:661–666. doi: 10.1016/s0896-6273(00)80726-6. [The author provides an insightful review of the experimental evidence for a role of postsynaptic action potentials in the induction of long-term depression and long-term potentiation, and the importance of the temporal order of pre- and postsynaptic activity for the sign of synaptic change] [DOI] [PubMed] [Google Scholar]
  • 36•.Johnston D, Hoffman DA, Colbert CM, Magee JC. Regulation of back-propagating action potentials in hippocampal neurons. Curr Opin Neurobiol. 1999;9:288–292. doi: 10.1016/s0959-4388(99)80042-7. [This is a review of recent data on the modulation of dendritic Na+ and K+ channels involved in regulating back-propagation of action potentials in hippocampal neurons. The authors emphasise a role for protein kinases in this regulation] [DOI] [PubMed] [Google Scholar]
  • 37•.Koester HJ, Sakmann B. Calcium dynamics in single spines during coincident pre- and postsynaptic activity depend on relative timing of back-propagating action potentials and subthreshold excitatory postsynaptic potentials. Proc Natl Acad Sci USA. 1998;95:9596–9601. doi: 10.1073/pnas.95.16.9596. [Spinous Ca2+ transients evoked by a single backpropagating action potential or subthreshold EPSP were of comparable amplitude. The Ca2+ transient evoked by pairing AP and EPSP showed both supra- and sublinear summation depending on the relative timing of AP and EPSP. The Ca2+ transient was larger if the AP followed the EPSP than if it preceded it] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38•.Schiller J, Schiller Y, Clapham DE. NMDA receptors amplify calcium influx into dendritic spines during associative pre- and postsynaptic activation. Nat Neurosci. 1998;1:114–118. doi: 10.1038/363. [Focal flash photolysis of caged glutamate produced spinous Ca2+ transients primarily via voltage-dependent Ca2+ channels. However, when glutamate was paired with postsynaptic action potentials, the NMDA-receptor-dependent component was selectively amplified] [DOI] [PubMed] [Google Scholar]
  • 39•.Emptage N, Bliss TV, Fine A. Single synaptic events evoke NMDA receptor-mediated release of calcium from internal stores in hippocampal dendritic spines. Neuron. 1999;22:115–124. doi: 10.1016/s0896-6273(00)80683-2. [The authors describe evidence from hippocampal slice cultures that synaptic NMDA receptor activation can trigger Ca2+-induced Ca2+ release in individual spines] [DOI] [PubMed] [Google Scholar]
  • 40•.Mainen ZF, Malinow R, Svoboda K. Synaptic calcium transients in single spines indicate that NMDA receptors are not saturated. Nature. 1999;399:151–155. doi: 10.1038/20187. [The authors provide evidence from acute hippocampal slices that the Ca2+ transient evoked by synaptic NMDA receptor activation originates primarily from the NMDAR channel and that Ca2+-induced Ca2+ release does not contribute to spine Ca2+ transients] [DOI] [PubMed] [Google Scholar]
  • 41.Markram H, Helm PJ, Sakmann B. Dendritic calcium transients evoked by single backpropagating action potentials in rat neocortical pyramidal neurons. J Physiol. 1995;485:1–20. doi: 10.1113/jphysiol.1995.sp020708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42•.Stiles JS, Bartol RM, Salpeter MM, Salpeter EE, Sejnowski TJ. Synaptic variability: new insights from reconstructions and Monte Carlo simulations with MCell. In: Cowan M, Davies K, editors. Synapses. Johns Hopkins University Press; Baltimore: 2000. [A general overview of MCell, a computer program that uses Monte Carlo techniques to simulate subcellular signaling in cells] [Google Scholar]
  • 43.Franks KM, Bartol TM, Egelman DM, Poo MM, Sejnowski TJ. Simulated dendritic influx of calcium ions through voltage- and ligand-gated channels using MCell. Soc Neurosci Abstr. 1999;25:1989. [Google Scholar]
  • 44••.Thomas MJ, Watabe AM, Moody TD, Makhinson M, O'Dell TJ. Postsynaptic complex spike bursting enables the induction of LTP by theta frequency synaptic stimulation. J Neurosci. 1998;18:7118–7126. doi: 10.1523/JNEUROSCI.18-18-07118.1998. [Using 5 Hz afferent stimulation in the CA1 area of hippocampal slices prepared from 4–7 week old mice, the authors describe a correlation between postsynaptic complex spike bursting and the induction of LTP. This finding indicates that postsynaptic bursting might have an important role in synaptic plasticity induced by physiologically relevant synaptic activity] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45••.Pike FG, Meredith RM, Olding AW, Paulsen O. Postsynaptic bursting is essential for ‘Hebbian’ induction of associative long-term potentiation at excitatory synapses in rat hippocampus. J Physiol. 1999;518:571–576. doi: 10.1111/j.1469-7793.1999.0571p.x. [In the CA1 area of hippocampal slices prepared from young adult rat, the authors report that postsynaptic burst firing is necessary for the induction of LTP using an experimental paradigm of pairing pre- and postsynaptic evoked activity. This finding supports the conclusion made in [44••] that postsynaptic burst firing is important for the induction of LTP by physiologically relevant synaptic activity in the adult hippocampus] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Mayford M, Wang J, Kandel ER, O'Dell TJ. CaMKII regulates the frequency-response function of hippocampal synapses for the production of both LTD and LTP. Cell. 1995;81:891–904. doi: 10.1016/0092-8674(95)90009-8. [DOI] [PubMed] [Google Scholar]
  • 47.Mayford M, Bach ME, Huang YY, Wang L, Hawkins RD, Kandel ER. Control of memory formation through regulated expression of a CaMKII transgene. Science. 1996;274:1678–1683. doi: 10.1126/science.274.5293.1678. [DOI] [PubMed] [Google Scholar]
  • 48•.Migaud M, Charlesworth P, Dempster M, Webster LC, Watabe AM, Makhinson M, He Y, Ramsay MF, Morris RGM, Morrison JH, O'Dell TJ, Grant SG. Enhanced long-term potentiation and impaired learning in mice with mutant postsynaptic density-95 protein. Nature. 1998;396:433–439. doi: 10.1038/24790. [The frequency threshold for induction of LTP was shifted towards lower frequencies in mice with mutant PSD-95. Surprisingly, this enhancement of LTP was associated with impaired learning in a water maze task] [DOI] [PubMed] [Google Scholar]
  • 49.Bienenstock EL, Cooper LN, Munro PW. Theory for the development of neuron selectivity: orientation specificity and binocular interaction in visual cortex. J Neurosci. 1982;2:32–48. doi: 10.1523/JNEUROSCI.02-01-00032.1982. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50•.Wang H, Wagner JJ. Priming-induced shift in synaptic plasticity in the rat hippocampus. J Neurophysiol. 1999;82:2024–2028. doi: 10.1152/jn.1999.82.4.2024. [Using field recordings from the CA1 area of hippocampal slices, the authors induced synaptic plasticity by a train of 600 pulses at a frequency between 3 and 100 Hz. They observed a >5-fold rightward shift in the frequency threshold between long-term depression and long-term potentiation after priming of the Schaffer collateral input with two sets of three high-frequency trains (100 Hz, 1 s). Since the effect of priming was expressed heterosynaptically as well as homosynaptically, this result suggests that the history of the neuron is important for the plastic properties of the synapses onto that neuron (metaplasticity)] [DOI] [PubMed] [Google Scholar]
  • 51••.Larkum ME, Zhu JJ, Sakmann B. A new cellular mechanism for coupling inputs arriving at different cortical layers. Nature. 1999;398:338–341. doi: 10.1038/18686. [Burst-associated dendritic Ca2+ spikes in neocortical pyramidal neurons can be triggered by otherwise subthreshold distal EPSPs when occurring in conjunction with a backpropagating action potential. The lowest threshold for initiation of Ca2+ spikes was seen when the EPSP occurred 5 ms after the backpropagating action potential] [DOI] [PubMed] [Google Scholar]
  • 52.Golding NL, Jung Hy, Mickus T, Spruston N. Dendritic calcium spike initiation and repolarization are controlled by distinct potassium channel subtypes in CA1 pyramidal neurons. J Neurosci. 1999;19:8789–8798. doi: 10.1523/JNEUROSCI.19-20-08789.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Seamans JK, Gorelova NA, Yang CR. Contributions of voltage-gated Ca2+ channels in the proximal versus distal dendrites to synaptic integration in prefrontal cortical neurons. J Neurosci. 1997;17:5936–5948. doi: 10.1523/JNEUROSCI.17-15-05936.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54•.Isomura Y, Kato N. Action potential-induced dendritic calcium dynamics correlated with synaptic plasticity in developing hippocampal pyramidal cells. J Neurophysiol. 1999;82:1993–1999. doi: 10.1152/jn.1999.82.4.1993. [The authors report that a developmental enhancement of action-potential-induced Ca2+ influx mediated by high-threshold voltage-dependent Ca2+ channels correlates with the induction of LTP by a theta-burst stimulation paradigm] [DOI] [PubMed] [Google Scholar]
  • 55•.Kudrimoti HS, Barnes CA, McNaughton BL. Reactivation of hippocampal cell assemblies: effects of behavioral state, experience, and EEG dynamics. J Neurosci. 1999;19:4090–4101. doi: 10.1523/JNEUROSCI.19-10-04090.1999. [In this paper, reactivation of experience-specific firing rate correlations between pairs of pyramidal neurons was seen during sharp waves that occurred during subsequent slow-wave sleep and quiet wakefulness] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56••.Nadasdy Z, Hirase H, Czurko A, Csicsvari J, Buzsaki G. Replay and time compression of recurring spike sequences in the hippocampus. J Neurosci. 1999;19:9497–9507. doi: 10.1523/JNEUROSCI.19-21-09497.1999. [In this paper, experience-specific spike sequences observed in the behaving rat were ‘replayed’ at a faster time-scale during sharp waves seen during subsequent slow-wave sleep] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57•.King C, Henze DA, Leinekugel X, Buzsaki G. Hebbian modification of a hippocampal population pattern in the rat. J Physiol. 1999;521:159–167. doi: 10.1111/j.1469-7793.1999.00159.x. [The authors demonstrate that strong depolarisation of individual CA1 pyramidal neurons during sharp waves can increase the cellular responsiveness to subsequent sharp waves] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Fields RD, Eshete F, Stevens B, Itoh K. Action potential-dependent regulation of gene expression: temporal specificity in Ca2+, cAMP-responsive element binding proteins, and mitogen-activated protein kinase signaling. J Neurosci. 1997;17:7252–7266. doi: 10.1523/JNEUROSCI.17-19-07252.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Frey U, Morris RGM. Synaptic tagging and long-term potentiation. Nature. 1997;385:533–536. doi: 10.1038/385533a0. [DOI] [PubMed] [Google Scholar]
  • 60••.Frey U, Morris RGM. Synaptic tagging: implications for late maintenance of hippocampal long-term potentiation. Trends Neurosci. 1998;21:181–188. doi: 10.1016/s0166-2236(97)01189-2. [A review by the discoverers of the phenomenon which led to the synaptic tagging concept] [DOI] [PubMed] [Google Scholar]
  • 61.Dudek SM, Fields RD. Somatic action potentials are sufficient for rescue of tagged synapses. Soc Neurosci Abstr. 1998;24:1074. [Google Scholar]
  • 62.Steriade M, McCormick DA, Sejnowski TJ. Thalamocortical oscillations in the sleeping and aroused brain. Science. 1993;262:679–685. doi: 10.1126/science.8235588. [DOI] [PubMed] [Google Scholar]
  • 63.Sejnowski TJ. Sleep and memory. Curr Biol. 1995;5:832–834. doi: 10.1016/s0960-9822(95)00165-5. [DOI] [PubMed] [Google Scholar]
  • 64.Deisseroth K, Bito H, Tsien RW. Signaling from synapse to nucleus: postsynaptic CREB phosphorylation during multiple forms of hippocampal synaptic plasticity. Neuron. 1996;16:89–101. doi: 10.1016/s0896-6273(00)80026-4. [DOI] [PubMed] [Google Scholar]
  • 65.Mainen ZF, Joerges J, Huguenard JR, Sejnowski TJ. A model of spike initiation in neocortical pyramidal neurons. Neuron. 1995;15:1427–1443. doi: 10.1016/0896-6273(95)90020-9. [DOI] [PubMed] [Google Scholar]
  • 66.Montague PR, Sejnowski TJ. The predictive brain: temporal coincidence and temporal order in synaptic learning mechanisms. Learn Memory. 1994;1:1–33. [PubMed] [Google Scholar]
  • 67.Schultz W, Dayan P, Montague PR. A neural substrate of prediction and reward. Science. 1997;275:1593–1599. doi: 10.1126/science.275.5306.1593. [DOI] [PubMed] [Google Scholar]
  • 68.Sejnowski TJ. Statistical constraints on synaptic plasticity. J Theor Biol. 1977;69:385–389. doi: 10.1016/0022-5193(77)90146-1. [DOI] [PubMed] [Google Scholar]
  • 69•.Abbott LF, Song S. Temporally asymmetric Hebbian learning, spike timing and neuronal response variability. In: Kearns MS, Solla SA, Cohn DA, editors. Advances in Neural Information Processing Systems 11. MIT Press; Cambridge, MA: 1999. pp. 69–75. [The authors show that a recurrent network of spiking neurons with time-dependent plasticity achieves a balanced condition where the interspike intervals are irregular, similar to that observed in cortex] [Google Scholar]
  • 70•.Seung HS. Learning continuous attractors in recurrent networks. In: Kearns M, Jordan M, Solla S, editors. Advances in Neural Information Processing Systems 10. MIT Press; Cambridge, MA: 1998. pp. 654–660. [Uses the anti-Hebbian form of the temporally asymmetric learning rule in a recurrent network to form a stable neural integrator] [Google Scholar]
  • 71.Bell CC, Han VZ, Sugawara Y, Grant K. Synaptic plasticity in a cerebellum-like structure depends on temporal order. Nature. 1997;387:278–281. doi: 10.1038/387278a0. [DOI] [PubMed] [Google Scholar]
  • 72.Abbott LF, Munro P. Neural Information Processing Systems (NIPS). Workshop on spike timing and synaptic plasticity; Breckenridge, CO.. December 4, 1999; http://www.pitt.edu/~pwm/LTP_LTD_99/ [Google Scholar]
  • 73•.Sejnowski TJ. The book of Hebb. Neuron. 1999;24:773–776. doi: 10.1016/s0896-6273(00)81025-9. [The author reviews the motivation for the Hebb synapse and shows how it is related to recent theoretical ideas about cell assemblies] [DOI] [PubMed] [Google Scholar]

RESOURCES