Next Article in Journal
Amorphous Calcium Carbonate from Plants Can Promote Bone Growth in Growing Rats
Previous Article in Journal
Advances in the Development of Biofertilizers and Biostimulants from Microalgae
Previous Article in Special Issue
Morphological and Functional Principles Governing the Plasticity Reserve in the Cerebellum: The Cortico-Deep Cerebellar Nuclei Loop Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Cerebellar Intrinsic Neuronal Excitability, Synaptic Plasticity, and Perineuronal Nets in Eyeblink Conditioning

by
Bernard G. Schreurs
1,*,
Deidre E. O’Dell
2 and
Desheng Wang
1
1
Department of Neuroscience, West Virginia University, Morgantown, WV 26505, USA
2
Department of Biology, Earth and Environmental Sciences, Pennsylvania Western (PennWest) University, California, PA 15419, USA
*
Author to whom correspondence should be addressed.
Biology 2024, 13(3), 200; https://doi.org/10.3390/biology13030200
Submission received: 14 January 2024 / Revised: 29 February 2024 / Accepted: 19 March 2024 / Published: 21 March 2024
(This article belongs to the Special Issue Plasticity and Computation in Cerebellar Neurons and Microcircuits)

Abstract

:

Simple Summary

Eyeblink conditioning is a simple form of learning that has been used to study areas of the brain involved in how we learn new tasks and how we remember them. One area of the brain that is important for eyeblink conditioning is the cerebellum. Changes that take place in the cerebellum involve a number of neural processes, including changes in the connections between neurons, changes in a neuron’s excitability, and even changes in the matrix that surrounds these neurons. Here, we explore these different processes and how they interact with each other to form the building blocks of a basic form of learning. Understanding how learning and memory take place may help us solve the mystery of how we lose the ability to learn and remember in diseases like Alzheimer’s disease, and how to we remember too much in post-traumatic stress disorder.

Abstract

Evidence is strong that, in addition to fine motor control, there is an important role for the cerebellum in cognition and emotion. The deep nuclei of the mammalian cerebellum also contain the highest density of perineural nets—mesh-like structures that surround neurons—in the brain, and it appears there may be a connection between these nets and cognitive processes, particularly learning and memory. Here, we review how the cerebellum is involved in eyeblink conditioning—a particularly well-understood form of learning and memory—and focus on the role of perineuronal nets in intrinsic membrane excitability and synaptic plasticity that underlie eyeblink conditioning. We explore the development and role of perineuronal nets and the in vivo and in vitro evidence that manipulations of the perineuronal net in the deep cerebellar nuclei affect eyeblink conditioning. Together, these findings provide evidence of an important role for perineuronal net in learning and memory.

1. Introduction

There is a long history of studying learning and memory to understand how a large range of organisms, including humans, adapt to the demands of their environment. One particularly well-understood form of learning is classical conditioning, first described by Pavlov more than 100 years ago. The defining features of classical conditioning include the delivery of a relatively innocuous signal or warning followed, almost immediately, by a significant event. In the case of fear conditioning, a tone is usually followed by shock to the feet of a rat or mouse, or the fingers of a human. In the case of eyeblink conditioning, the same tone may be followed by a puff of air to the eye of a person, monkey, rabbit, rat, or mouse. The history of eyeblink conditioning began with the study of behavioral laws governing the acquisition, consolidation, and extinction of a conditioned response—closure of the eye during the tone and before the air puff. A growing interest in understanding not only the “how” but the “where” of eyeblink conditioning led to concerted efforts to search for the sites in the brain where learning and memory take place—the engram—and the particular role of the cerebellum. Although not universally accepted, there is evidence that the cerebellum is an engram for eyeblink conditioning. Research shows changes in synaptic and intrinsic membrane plasticity, as well as changes in perineuronal nets, are involved in the successful acquisition of eyeblink conditioning. We have previously reviewed evidence that eyeblink conditioning results in significant changes in intrinsic membrane excitability and synaptic plasticity in the cerebellum. Here, we review research showing that perineuronal nets surrounding principal neurons in the deep cerebellar nuclei (DCN) are involved in eyeblink conditioning and may mediate changes in intrinsic membrane excitability and synaptic plasticity.

2. Eyeblink Conditioning

Although head-fixed mice are the most recent research subjects to undergo eyeblink conditioning, Hilgard and colleagues reported eyeblink conditioning in the 1930s in dogs [1] and humans [2]. Considerable theoretical and experimental interest in human eyeblink conditioning continued into the 1960s [3], but conceptual and methodological issues dampened enthusiasm for human eyeblink research [4,5]. First, Gormezano and others raised conceptual and methodological concerns about the exclusion of subjects known as “voluntary responders”, who deliberately blinked to the tone or light during an eyeblink conditioning experiment [4]. Second, there was a growing frustration with an inability to analyze the brain of human subjects—sometimes considered a black box—to identify the biological basis of learning and memory. The latter limitation has since been overcome, in large part, by the use of imaging techniques, including positron emission tomography [6,7,8] and magnetic resonance imaging [9,10,11], that have identified the involvement of areas including the hippocampus, prefrontal cortex, and cerebellum in human eyeblink conditioning. These efforts to identify areas of the brain involved in human eyeblink conditioning were antedated in the 1980s by a shifting interest to search for the engram [12,13,14,15,16,17] in animal models of eyeblink conditioning [18].
The basic eyeblink conditioning paradigm involves presenting a conditioned stimulus (CS), such as a tone or light, that does not initially elicit an eyeblink response, and pairing it with an unconditioned stimulus (US), such as a puff of air or brief electrical pulse, near the eye that does elicit an eyeblink response (unconditioned response, UR). With repeated pairings of the two stimuli at specific durations and intervals, a conditioned eyeblink response (CR) can emerge. Importantly, explicitly unpaired presentations of the tone and air puff or shock to a different group of subjects to assess nonassociative contributors to responding, such as sensitization, show very low levels of responding to the CS [19,20,21]. In addition to the initial research with rabbits, interest in the use of rats to determine the neural substrates of eyeblink conditioning increased when Skelton overcame the limitations of restraint normally required for eyeblink conditioning in rats [22] by developing a procedure for recording eyelid EMG activity in freely moving animals [23,24,25,26,27,28]. Figure 1 shows an unpublished waterfall plot of responding in a freely moving adult rat given paired CS–US presentations. More recently, mouse eyeblink conditioning has been explored with several different recording techniques using head-fixed mice that walk on a movable surface, such as a ball or cylinder [29,30,31,32,33,34,35,36].

3. The Role of the Cerebellum in Eyeblink Conditioning

Early results in the search for the engram for eyeblink conditioning in rabbits documented the involvement of the hippocampus [37,38], and later findings described the more crucial role played by the cerebellum [39,40,41,42,43]. Lesion, inactivation, and recording studies extensively reviewed elsewhere [44,45,46,47,48,49,50] identified the deep cerebellar nuclei (DCN), particularly the anterior interpositus nucleus (AIN), as being important for eyeblink conditioning [36,50,51,52,53,54,55,56,57,58,59]. This was not a universally accepted position [48,60,61,62]. For example, in a consensus paper, Perciavalle et al. [63] provided evidence from rats, cats, and rabbits, among others, that the AIN participated in the timing and performance of ongoing conditioned eyeblinks but did not generate or initiate the conditioned response. Earlier, Welsh and Harvey showed that lesions of the rabbit AIN resulted in motor performance deficits in response timing and amplitude in both conditioned and unconditioned responses, but did not eliminate conditioned responses [60]. A similar debate has occurred about the role of the cerebellar cortex in eyeblink conditioning, particularly lobule HVI, which Yeo first showed was important in rabbit eyeblink conditioning [43,64,65]. As with the role of the AIN, the role of HVI in mice, rats, rabbits, and humans has been shown to be important [10,11,29,51,64,65,66,67,68] but not the sole site [69,70] for eyeblink conditioning. As with many controversies in the field, the devil is in the details, and the range of stimuli (e.g., tones versus lights, shock versus air puff), site and extent of the lesions (unilateral versus bilateral, reversible versus permanent), the conditioning parameters (interstimulus interval, intertrial interval, number of trials), method of restraint (head-fixed, cloth bag, solid restrainer, freely moving), modes of assessment (EMG, movement potentiometer, non-invasive reflectance, inductive coils, magnetic field effects), not to mention different species (mouse, rat, rabbit, cat, human) used to assess eyeblink conditioning, may all be important. All of these issues notwithstanding, research across these parameters, methods, and assessment modes with a very diverse range of species, including goldfish [71], turtles [72], rats [73], mice [74], guinea pigs [75,76,77], rabbits [41,60,70], and cats [78,79,80], as well as humans [8,81], suggests an important role for the cerebellum in eyeblink conditioning.
The question then becomes: what is the nature of the changes that take place in the cerebellum as a result of eyeblink conditioning? To understand the evidence, we first need to review the basic cerebellar circuitry underlying eyeblink conditioning. Figure 2 from [82] is a simplified illustration of some of the essential cerebellar circuitry involved in eyeblink conditioning. The figure shows that sensory inputs from the tone give rise to mossy fibers (MF) that synapse onto neurons in the AIN, and onto granule cells that give rise to parallel fibers that synapse onto Purkinje cell dendrites. Sensory inputs from the air puff give rise to climbing fibers (CF) from the inferior olive that also synapse onto neurons in the AIN, and then encircle the dendrites of a Purkinje cell.
Evidence of eyeblink-specific changes in the rabbit AIN was first reported as altered extracellular activity recorded with metal electrodes [39,41], and later also reported in the rat AIN [83]. Subsequently, evidence of conditioning-specific changes in the AIN began to emerge from electron microscopy, but those results remain somewhat confusing. On the one hand, there were no overall changes in synaptic number, shape, or perforations in excitatory neurons in the AIN of the DCN, although there was a significant increase in the length of excitatory synapses after rabbit eyeblink conditioning [84]. No distinction was made between mossy fibers and climbing fiber inputs. On the other hand, there was an increase in the number of excitatory synapses, but not inhibitory synapses in the rat DCN after eyeblink conditioning [85]. More recent work by Broersen et al. [86] has expanded on the Kleim et al. [85] study by showing an increase in the number of mossy fiber excitatory terminals, identified by VGlut1 labeling, in presumed projection neurons in the AIN of the mouse DCN as a result of eyeblink conditioning. In contrast to the Kleim study, Broersen et al. also showed an increase in the number of Purkinje cell inhibitory terminals to presumed AIN projection neurons as a result of eyeblink conditioning in the mouse. If the DCN is so important to eyeblink conditioning, it is difficult to reconcile these differences by simply pointing to the different species (rabbit versus rat versus mouse) or the different eyeblink conditioning procedures (tone versus light, air puff versus shock) that were employed. Perhaps it was an improvement in the sophistication of the techniques used to make the measurements. On the other hand, the changes may not have occurred exclusively in neurons involved in eyelid conditioning. For example, in addition to eyeblinks, Broersen et al. noted whole-body movements evoked by the conditioned stimulus. This idea may be supported by our finding that recordings from projection neurons in the juvenile rat AIN, labeled with a pseudorabies transsynaptic virus injected into the eyelid, showed increases in intrinsic membrane excitability following eyeblink conditioning in both labeled and unlabeled neurons [87]. These increases in excitability were significantly higher than in rats given unpaired presentations of the tone and shock.

4. Intrinsic Membrane Excitability Involved in Eyeblink Conditioning

A substantial amount of work has been performed to understand changes in intrinsic membrane excitability—changes in the likelihood of generating an action potential—and how these changes are involved in eyeblink conditioning. Changes in intrinsic membrane excitability are changes in the movement of ions across the membrane and come about by the insertion, removal, or modification of sodium, calcium, potassium, or non-selective cation channels [88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111]. These changes can occur as an increase or decrease in input resistance, membrane potential, afterhyperpolarization, spike threshold, spike frequency, or a combination of these properties. Intrinsic membrane excitability changes—reviewed extensively elsewhere [44,99,112,113,114,115,116,117]—have been found across a range of structures, species, and learning paradigms. Of relevance to this review are the experience-dependent changes in membrane properties found in the cerebellum [34,44,87,88,90,95,101,111,114,117,118,119,120,121,122,123,124,125], particularly eyeblink conditioning [34,44,87,126]. In a series of mouse eyeblink conditioning experiments, Titley, Hansel and colleagues examined changes in Purkinje cell membrane excitability [34] and found that increases in excitability were mediated by a Purkinje cell-specific calcium-activated K+ channel (SK2) [126]. We have also reported learning-specific changes in intrinsic membrane excitability in Purkinje cell dendrites, but as a function of rabbit eyeblink conditioning [127,128,129]. We were able to show that 4-AP—which blocks a transient voltage-dependent potassium channel at low concentrations—reduced afterhyperpolarization-mediated changes in excitability [127,130] and concluded that the observed eyeblink conditioning-specific increase in dendritic excitability was a function of changes in an IA-like potassium current—a finding consistent with changes in hippocampal CA1 pyramidal cells and interneurons as a function of eyeblink conditioning in the rabbit [131,132]. We have also reported an increase in intrinsic membrane excitability, measured as a reduction in the afterhyperpolarization of projection neurons in the AIN as a function of eyeblink conditioning in juvenile rats [87]. Importantly, dendritic excitability has been shown to be increased by the trafficking of voltage-dependent potassium channels [92].

5. Synaptic Plasticity in the Cerebellum Involved in Eyeblink Conditioning

Synaptic plasticity in the cerebellar cortex takes the form of long-term depression (LTD) and long-term potentiation (LTP)—phenomena that have been reviewed elsewhere [133,134,135,136,137,138,139]. LTD and LTP in the deep cerebellar nuclei have also been studied extensively [140,141,142,143,144], particularly in slice work by Raman and colleagues [145,146,147,148,149]. In the cerebellar cortex, stimulating climbing fibers and parallel fibers that synapse onto Purkinje cells reduces parallel fiber synaptic potentials—first proposed by Marr [150] and Albus [151] and later confirmed by Ito [152,153] and Gilbert and Thach [154]—has become a benchmark for studying the molecular underpinnings of synaptic plasticity. In the DCN, stimulation of white matter, which activates mossy fiber collaterals, has been shown to result in LTP of large, fast-spiking excitatory neurons, but only in the presence of hyperpolarization that mimics Purkinje cell inhibition [147,155]. Stimulation of Purkinje cell inhibitory inputs to these same large, fast-spiking excitatory neurons has been shown to result in either LTP or LTD depending on the stimulation parameters [143,144,156,157,158].
Although Ito first used cerebellar LTD to explain experience-dependent adaptation of the vestibulo-ocular reflex [159], it has also been used to explain eyeblink conditioning [138,160,161,162,163,164,165]. As we saw in Figure 2, information about the tone and air puff used in eyeblink conditioning reaches the cerebellum through mossy fibers and climbing fibers that synapse at the DCN and Purkinje cells [166,167,168]. In the case of delay eyeblink conditioning, where the tone and air puff overlap and co-terminate, activation of parallel and climbing fibers occurs together during the overlap, which could result in long-term depression of Purkinje cell inhibition of the DCN and long-term potentiation of mossy fiber excitation. Significantly, Broersen et al. [86] found an increase in the number of Purkinje cell inhibitory terminals and mossy fiber excitatory terminals in the AIN of the mouse DCN that may mediate LTD and LTP, respectively, as a result of eyeblink conditioning. As with the contrasting views about the role of the cerebellum in eyeblink conditioning described above, there are also contrasting views on the role of cerebellar synaptic plasticity in eyeblink conditioning. Some argue that it is necessary for eyeblink conditioning [169,170,171,172], and others have shown that eyeblink conditioning can occur in the absence of synaptic plasticity [161,173,174,175]. There are also more nuanced positions suggesting that synaptic plasticity is one of several forms of plasticity involved in eyeblink conditioning [44,114,117,174].
Although much of the debate about the role of synaptic plasticity in eyeblink conditioning has been based on delay conditioning where stimuli overlap, eyeblink conditioning in many species, including rats, rabbits, and humans, can be achieved with a trace conditioning paradigm where the tone ends 250–500 milliseconds before the air puff, creating a gap or “trace” between the stimuli [28,176,177,178,179,180,181,182,183]. As reviewed elsewhere [44], delays in transmission along the mossy/parallel fiber circuit have been proposed that ensure sensory inputs from the tone and air puff arrive in the cerebellum at the same time [184]. However, auditory transmission rates [185] and direct electrical stimulation that substitutes for tone and air puff all suggest there is no significant delay in activation of mossy fibers or climbing fibers [186]. Others have found that the medial prefrontal cortex may bridge the trace and then activate mossy fibers that are contiguous with a climbing fiber input [187,188,189]. This is, in essence, a serial compound of two sequential mossy fiber inputs (directly from the brainstem auditory system and then indirectly from the prefrontal cortex) followed by a climbing fiber input. Kehoe and others have explored serial compound conditioning extensively, and there is very good evidence that serial compounds of two stimuli allow animals to bridge temporal gaps of the order of seconds—gaps that are otherwise too long to support eyeblink conditioning [190,191,192,193,194]. However, behavioral studies of serial compounds have invariably used two different stimuli—usually a tone and a light—and it is not clear that if the same stimuli were presented serially, the temporal gap would be bridged. When two identical stimuli have been used in sequence, rabbit and rat eyelid conditioning has been achieved, but the stimuli were both periorbital shocks that elicited an eyelid response [195,196]. It might be too simple to suggest that these shocks only elicited climbing fiber inputs to the cerebellum, although evidence suggests that climbing fiber activation can induce plasticity [197,198]. Shock has been described as an “inadequate” stimulus because direct electrical stimulation of nerve fibers in the region are qualitatively different from the sensory transduction that is required to detect light, sound, and touch. Nevertheless, shock elicits sensations like buzzing and pain that could reach the cerebellum along mossy fibers. In other words, presentation of two serial shocks would consist of both MF and CF inputs to the cerebellum that could alter synaptic properties leading to eyeblink conditioning [195].

6. Membrane Excitability and Synaptic Plasticity in Eyeblink Conditioning

The question is: how do distant changes in synaptic plasticity at specific dendrites translate to changes in the output of the cell [14,44,93,99,113,114,123,199]? This is particularly true in Purkinje cells where the dendritic tree is extensive, and there are thousands of synaptic connections from parallel fibers, as well as a smaller number of inhibitory connections from molecular layer interneurons. If, as documented elsewhere [44,127,129], there are both membrane and synaptic conditioning-specific changes following eyeblink conditioning, what is the sequence of those changes? Does synaptic plasticity occur first, followed by membrane plasticity, or vice versa? If membrane excitability increases occur first, then any changes at specific synapses in the dendrites—which would be considered to be more subtle because of the size of the dendritic tree—would presumably be amplified. Conversely, if synaptic plasticity changes occurred first, they would have to wait for changes in membrane excitability to be effective in altering a neuron’s output. What does this mean for behavior? If cells in the deep cerebellar nuclei are more excitable, the same synaptic input from MFs would cause AIN neurons, which initially only fired to the input provided by both the MFs and CF collaterals, to fire more readily, leading to an increased likelihood of a CR when only the MF collaterals have input to the AIN, which is what we see on tone-alone test trials. This would then be amplified by synaptic plasticity—the decreased inhibition of deep cerebellar nuclei from Purkinje cell axons and increased excitation from mossy fiber inputs already described above [86]. Although the evidence for the involvement of both synaptic plasticity and membrane excitability in eyeblink conditioning reviewed so far is persuasive, this may only be part of the story, because there is increasing evidence that the perineuronal nets surrounding the deep cerebellar nuclei may also play a role in eyeblink conditioning [200,201,202,203].

7. The Perineuronal Net

Cajal may have been the first to observe the perineuronal net (PNN) in the cerebellum, but Golgi appreciated that the PNN was more than a tissue processing artifact and is credited with identifying the PNN as a structure and describing it in detail [204,205]. The original concept of communication in the brain comprising synaptic connections between neurons at a pre- and post-synaptic interface, first proposed by Cajal [168,206], has since been expanded to include the vital role of glia that together form the tripartite synapse [207,208,209]. More recently, the importance of the PNN surrounding the soma and proximal dendrites of a neuron, and being actively involved in modulating communication, has prompted the notion of a tetrapartite structure [210,211,212,213].
The PNN is a form of extracellular matrix that forms a reticular structure around several different classes of neurons in the brain, particularly fast-spiking neurons [204,214,215,216,217,218,219,220], including projection neurons in the DCN that can fire at rates in excess of 100 Hz [221,222]. In fact, the PNN is found covering more cells in the DCN than any other part of the brain [223,224]. The PNN is composed of hyaluronic acid, link proteins, chondroitin sulfate proteoglycans (CSPGs), and tenascin-R that assemble into a dense, lattice-like sheet that can be disrupted by chondroitinase ABC (ChABC), an enzyme that degrades the glycosaminoglycan side chains of chondroitin sulfate proteoglycans. Development of the PNN has correlated with critical periods of plasticity [225,226,227,228,229,230]. The Bruckner group identified stages of PNN maturation across different regions of the postnatal rat brain from P0 [231] as well as by Ye and Miao, who studied PNN development in the postnatal mouse visual cortex from P10 to P42 [232]. We have reported development of the PNN in the rat DCN, where the PNN does not fully assemble into its lattice-like structure until the end of the third week of post-natal development [201]. Figure 3, adapted from [201], shows the development of the PNN in the rat DCN—determined by the labeling of CSPGs with WFA (Wisteria floribunda agglutinin). The figure shows that the PNN surrounding neurons in the DCN does not fully develop in rat pups until after post-natal day 18 (P18). Interestingly, this is also around the time when electrical properties of neurons in the rat DCN mature with increases in the amplitude of the afterhyperpolarization, a prolonged interval between the first and second evoked action potential, and an increase in afterhyperpolarization amplitude for hyperpolarization-induced rebound spikes [233]. This is also when climbing fiber pruning nears completion [234] and rats are first able to acquire eyeblink conditioning to either tones or lights paired with shock [201,235,236,237,238,239], although they can acquire conditioned responses at early time points if other stimuli are used, including two shocks [196] or direct brain stimulation [240].
The role of the PNN in eyeblink conditioning has been explored by several groups [200,201,202,203]. Hirono et al. used enzymatic digestion of the PNN with chondroitinase ABC (ChABC) in cerebellar slices of the DCN and found a decrease in Purkinje cell inhibitory postsynaptic currents, as well as higher terminal levels of eyeblink conditioning in head-fixed mice treated with ChABC compared to mice infused with the vehicle [203]. Carulli et al. used a lentiviral approach to release ChABC into the mouse DCN and showed a reduction in spontaneous activity of DCN neurons that may have been due to increased Purkinje cell inhibitory inputs and decreased mossy fiber excitatory inputs. They suggested that this could explain the enhanced plasticity in the DCN during the acquisition of eyeblink conditioning [202,241]. We recently showed that in vivo degradation of the PNN by ChABC using indwelling cannulae resulted in significant reductions in freely moving rat eyeblink conditioning amplitude and area compared to saline-infused controls, but did not affect conditioned or unconditioned response frequency [201]. Figure 4 shows an example of PNN digestion in the left DCN (ChABC) compared to a control infusion on the right (saline) four days following infusion. Although there was a 50% reduction in the percentage of neurons with WFA labeling, digestion in the DCN was not complete, and there were still neurons with WFA reactivity throughout the structure. The remaining neurons with WFA labeling may explain why there were significant changes in the amplitude and area of CRs without a reduction in the frequency of responding. We next conducted an in vitro experiment in which slices of the cerebellum were incubated with ChABC, and found the AIN had fewer WFA-positive neurons (41.98% ± 4.75) compared to the AIN in slices incubated with the vehicle (98.71% ± 0.38), p < 0.001. Neurons exposed to ChABC required more current to fire an action potential (AP) and had a longer latency to evoke an AP compared to cells in the vehicle group. AIN neurons exposed to ChABC also showed a longer inter-spike interval and had a larger afterhyperpolarization amplitude, shown in Figure 5. There also appeared to be a more robust digestion in our in vitro condition compared to our in vivo study. Interestingly, there were no differences in the membrane potential or input resistance. These results suggest that digestion of PNN with ChABC in acute AIN slices decreased the intrinsic excitability of large excitatory neurons without affecting other membrane properties. Although we saw decreased excitability in fast-spiking projection neurons in DCN in vitro, Hayani et al. saw no changes in mouse hippocampal fast-spiking interneurons or principal cells after treatment with ChABC in vitro [242]. In a comprehensive review of electrophysiological consequences of PNN modification, Wingert and Sorg (2018) concluded that removing PNNs has an impact on the synaptic and membrane properties of fast-spiking interneurons, but less so on principal neurons [241,242].
There has been considerable discussion about the role of the PNN in synaptic plasticity. In examining perineuronal nets of the adult rat cerebellum, Carulli et al. stated that PNNs have “holes at the sites of synaptic contacts” [243]. The concept of holes for synaptic contacts onto neurons in rat deep cerebellar nuclei was suggested by Lafraga et al., who noted the PNN had “holes for the synaptic boutons.” [244]—an idea proposed much earlier by Schwartz [245]. Tsien reflected on the function of the PNN by suggesting long-term memories were stored in the pattern made by these holes [246]. In a recent review, Rudolph et al. noted PNNs restrict new synapses from being produced and old synapses from being pruned, which is thought to regulate neuronal plasticity [247]. This view was included among those described by Celio et al., who also provided evidence that the PNN may maintain cellular relationships, concentrate growth factors, generate an ion-buffering microenvironment, prevent extracellular space occlusion, and form a link with the intracellular cytoskeleton [205]. Carulli et al. suggested that PNNs are strategically positioned to influence the development and stabilization of synaptic connections [202]. Frischknecht et al. showed that PNN removal facilitated AMPA receptor movement across the membrane, whereas NMDA receptors did not move with PNN removal [248].
There has been much less discussion about the role of the PNN in intrinsic membrane excitability. Theorizing about the function of the PNN includes regulating the localization of ion channels [249,250], binding cations [251], gating ion channels [252], anchoring ion channels, ion exchangers, and ion transporters in the plasma membrane, as well as reducing membrane capacitance by acting as an electrostatic insulator [253]. PNNs may act as local buffers of sodium and potassium ions in the extracellular space to ensure rapid ion transport [254,255]. In an extension of that idea, Morawski et al. suggested the PNN contains anionic binding sites that trap extracellular calcium, potassium, and sodium that can be mobilized in the service of the demands of fast-spiking neurons [256]. If PNNs act as an anchor for ion channels, it follows that disrupting the PNN may make ion channels more able to be inserted or removed. This could explain the decreases in membrane excitability we observed by disrupting the PNN with ChABC—an increase in voltage-dependent potassium channels in the membrane. In measuring the effects of the PNN on membrane properties, Frischknecht et al. showed that PNN removal did not affect resting potential, AP amplitude, or width, but they did not measure other indices of membrane excitability [248]. We also did not see changes in resting membrane potential or action potential amplitude as a function of treating the PNN with ChABC, but we did observe alterations in other measures of membrane excitability, particularly the amplitude of the afterhyperpolarization [201], which may have resulted from the insertion of voltage-dependent potassium channels.
Taken together, the data suggest the PNN may control cellular and synaptic forms of plasticity by regulating the localization of ion channels, particularly potassium channels, and receptors, particularly AMPA receptors [92,241,249]. It has even been suggested that these functions may be the result of modification of the PNN by microglia [225] through the manipulation of proteases and phagocytosis [212]. The absence of local microglia through experimental depletion has been shown to enhance PNN deposition and density, in addition to affecting synaptic number [212]. As we have noted, insights into the role of the PNN have also come from direct enzymatic and genetic manipulation of the PNN. A recent review by Fawcett et al. described the effects of modulating synaptic function by genetically or enzymatically perturbing the PNN and the potential effects on a large range of learning and memory, including eyeblink conditioning [257]. We suggest that perturbing the PNN may also have consequences for intrinsic membrane excitability—another form of plasticity that underlies eyeblink conditioning.

8. Conclusions

Eyeblink conditioning results in significant changes in intrinsic membrane excitability and synaptic plasticity in the cerebellum in a number of species. As in other preparations, there is a growing consensus that both intrinsic membrane excitability and synaptic plasticity are required for eyeblink conditioning [88,99,101,114,117,121,122,123,131,132,199,258,259,260,261,262,263]. More recently, there is a growing consensus that perineuronal nets are also involved in learning and memory across a range of paradigms [210,264,265,266,267,268,269,270], including fear conditioning [271,272,273,274,275,276,277,278] and eyeblink conditioning [86,200,201,202,203,257]. Together, these findings provide evidence for the combined role of intrinsic membrane excitability, synaptic plasticity, and the perineuronal net in eyeblink conditioning. Evidence is growing that the perineuronal net may alter learning and memory by regulating intrinsic membrane excitability and synaptic plasticity.

Author Contributions

Writing—original draft preparation, B.G.S.; writing—review and editing, B.G.S., D.E.O. and D.W.; visualization, B.G.S. and D.E.O.; funding acquisition, B.G.S. and D.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the NINDS intramural research program, NIH Grant NS094009, a bridge grant from West Virginia University Health Sciences Research Office, and Rockefeller Neuroscience Institute funds. PRV-152 was a kind gift from Lynn W. Enquist, who was funded by NIH Grant P40 OD010996. BGS and DW are supported in part by NIH grant HD099338 and DO was supported by NIH grants T32 GM081741 and HD099338. The contents of, and the views and opinions expressed in this article, are solely the responsibility of the authors and do not necessarily represent the official views of the NIH.

Acknowledgments

The authors are grateful to the graduate students and postdoctoral fellows, without whom this research would not have been possible.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hilgard, E.R.; Marquis, D.G. Acquisition, extinction, and retention of conditioned responses to light in dogs. J. Comp. Psychol. 1935, 19, 29–58. [Google Scholar] [CrossRef]
  2. Hilgard, E.R.; Campbell, A.A. The course of acquisition and retention of conditioned eyelid responses in man. J. Exp. Psychol. 1936, 19, 227–247. [Google Scholar] [CrossRef]
  3. Spence, K.W.; Spence, J.T. Sex and anxiety difference in eyelid conditioning. Psychol. Bull. 1966, 65, 137–142. [Google Scholar] [CrossRef] [PubMed]
  4. Coleman, S.R.; Webster, S. The problem of volition and the conditioned reflex. Part II: Voluntary-responding subjects, 1951–1980. Behaviorism 1988, 16, 17–49. [Google Scholar]
  5. Coleman, S.R. The problem of volition and the conditioned reflex. Part I: Conceptual background, 1900–1940. Behaviorism 1985, 13, 99–124. [Google Scholar]
  6. Blaxton, T.A.; Zeffiro, T.A.; Gabrieli, J.D.E.; Bookheimer, S.Y.; Carrillo, M.C.; Theodore, W.H.; Disterhoft, J.F. Functional mapping of human learning: A positron emission tomography activation study of eyeblink conditioning. J. Neurosci. 1996, 16, 4032–4040. [Google Scholar] [CrossRef] [PubMed]
  7. Logan, C.G.; Grafton, S.T. Functional anatomy of human eyeblink conditioning determined with regional cerebral glucose metabolism and positron-emission tomography. Proc. Natl. Acad. Sci. USA 1995, 92, 7500–7504. [Google Scholar] [CrossRef]
  8. Molchan, S.E.; Sunderland, T.; McIntosh, A.R.; Herscovitch, P.; Schreurs, B.G. A functional anatomical study of associative learning in humans. Proc. Natl. Acad. Sci. USA 1994, 91, 8122–8126. [Google Scholar] [CrossRef]
  9. Kent, J.S.; Kim, D.J.; Newman, S.D.; Bolbecker, A.R.; O’Donnell, B.F.; Hetrick, W.P. Investigating cerebellar neural function in schizophrenia using delay eyeblink conditioning: A pilot fMRI study. Psychiatry Res. Neuroimaging 2020, 304, 111133. [Google Scholar] [CrossRef]
  10. Dimitrova, A.; Weber, J.; Maschke, M.; Elles, H.-G.; Kolb, F.P.; Forsting, M.; Diener, H.-C.; Timmann, D. Eyeblink-related areas in human cerebellum as shown by fMRI. Hum. Brain Mapp. 2002, 17, 100–115. [Google Scholar] [CrossRef]
  11. Ramnani, N.; Toni, I.; Josephs, O.; Ashburner, J.; Passingham, R.E. Learning- and expectation-related changes in the human brain during motor learning. J. Neurophysiol. 2000, 84, 3026–3035. [Google Scholar] [CrossRef] [PubMed]
  12. Han, D.H.; Park, P.; Choi, D.I.; Bliss, T.V.P.; Kaang, B.K. The essence of the engram: Cellular or synaptic? Semin. Cell Dev. Biol. 2021, 125, 122–135. [Google Scholar] [CrossRef] [PubMed]
  13. Brodt, S.; Gais, S. Memory engrams in the neocortex. Neuroscientist 2021, 27, 427–444. [Google Scholar] [CrossRef]
  14. Josselyn, S.A.; Tonegawa, S. Memory engrams: Recalling the past and imagining the future. Science 2020, 367, eaaw4325. [Google Scholar] [CrossRef]
  15. Sakaguchi, M.; Hayashi, Y. Catching the engram: Strategies to examine the memory trace. Mol. Brain 2012, 5, 32. [Google Scholar] [CrossRef] [PubMed]
  16. Hubener, M.; Bonhoeffer, T. Searching for engrams. Neuron 2010, 67, 363–371. [Google Scholar] [CrossRef]
  17. Schafe, G.E.; Doyere, V.; LeDoux, J.E. Tracking the fear engram: The lateral amygdala is an essential locus of fear memory storage. J. Neurosci. 2005, 25, 10010–10015. [Google Scholar] [CrossRef]
  18. Steinmetz, J.E.; Woodruff-Pak, D.S. Animal models in eyeblink classical conditioning. In Eyeblink Classical Conditioning Volume 2: Animal Models; Woodruff-Pak, D.S., Steinmetz, J.E., Eds.; Kluwer Academic Publishers: Boston, MA, USA, 2000; pp. 1–15. [Google Scholar]
  19. Gormezano, I.; Kehoe, E.J.; Marshall, B.S. Twenty years of classical conditioning research with the rabbit. In Progress in Psychobiology and Physiological Psychology, 10th ed.; Sprague, J.M., Ed.; Academic Press: New York, NY, USA, 1983; pp. 197–275. [Google Scholar]
  20. Schneiderman, N.; Fuentes, I.; Gormezano, I. Acquisition and extinction of the classically conditioned eyelid response in the albino rabbit. Science 1962, 136, 650–652. [Google Scholar] [CrossRef]
  21. Gormezano, I.; Kehoe, E.J. Classical Conditioning: Some methodological-conceptual issues. In Handbook of Learning and Cognitive Processes (Volume 2): Conditioning and Behavior Theory; Estes, W.K., Ed.; Earlbaum: Hillsdale, NJ, USA, 1975; pp. 143–179. [Google Scholar]
  22. Schmajuk, N.A.; Christiansen, B.A. Eyeblink conditioning in rats. Physiol. Behav. 1990, 48, 755–758. [Google Scholar] [CrossRef]
  23. Weiss, C.; Thompson, R.F. The effects of age on eyeblink conditioning in the freely moving Fischer-344 rat. Neurobiol. Aging 1991, 12, 249–254. [Google Scholar] [CrossRef]
  24. Shors, T.J.; Servatius, R.J. The contribution of stressor intensity, duration, and context to the stress-induced facilitation of associative learning. Neurobiol. Learn. Mem. 1997, 67, 92–96. [Google Scholar] [CrossRef] [PubMed]
  25. Stanton, M.E.; Freeman, J.H.; Skelton, R.W. Eyeblink conditioning in the developing rat. Behav. Neurosci. 1992, 106, 657–665. [Google Scholar] [CrossRef] [PubMed]
  26. Knuttinen, M.-G.; Gamelli, A.E.; Weiss, C.; Power, J.M.; Disterhoft, J.F. Age-related effects on eyeblink conditioning in F344 X BN F1 hybrid rats. Neurobiol. Aging 2001, 22, 1–8. [Google Scholar] [CrossRef]
  27. Green, J.T.; Rogers, R.F.; Goodlett, C.R.; Steinmetz, J.E. Impairment in eyeblink classical conditioning in adult rats exposed to ethanol as neonates. Alcohol. Clin. Exp. Res. 2000, 24, 438–447. [Google Scholar] [CrossRef] [PubMed]
  28. Weiss, C.; Knuttinen, M.-G.; Power, J.M.; Patel, R.I.; O’Connor, M.S.; Disterhoft, J.F. Trace eyeblink conditioning in the freely moving rat: Optimizing the conditioning parameters. Behav. Neurosci. 1999, 113, 1100–1105. [Google Scholar] [CrossRef] [PubMed]
  29. Ten Brinke, M.M.; Boele, H.J.; Spanke, J.K.; Potters, J.W.; Kornysheva, K.; Wulff, P.; AC, I.J.; Koekkoek, S.K.; De Zeeuw, C.I. Evolving models of Pavlovian conditioning: Cerebellar cortical dynamics in awake behaving mice. Cell Rep. 2015, 13, 1977–1988. [Google Scholar] [CrossRef]
  30. Koekkoek, S.K.E.; Den Ouden, W.L.; Perry, G.; Highstein, S.M.; De Zeeuw, C.I. Monitoring kinetic and frequency-domain properties of eyelid responses in mice with magnetic distance measurement technique. J. Neurophysiol. 2002, 88, 2124–2133. [Google Scholar] [CrossRef]
  31. De Zeeuw, C.I.; Van Alphen, A.M.; Koekkoek, S.K.E.; Buharin, E.; Coesmans, M.P.H.; Morpurgo, M.M.; Van Der Burg, J. Recording eye movements in mice: A new approach to investigate the molecular basis of cerebellar control of motor learning and motor timing. Otolaryngol.-Head Neck Surg. 1998, 119, 193–203. [Google Scholar] [CrossRef]
  32. Miller, L.N.; Weiss, C.; Disterhoft, J.F. Learning-related changes in cellular activity within mouse dentate gyrus during trace eyeblink conditioning. Hippocampus 2022, 32, 776–794. [Google Scholar] [CrossRef]
  33. de Oude, N.L.; Hoebeek, F.E.; Ten Brinke, M.M.; de Zeeuw, C.I.; Boele, H.J. Pavlovian eyeblink conditioning is severely impaired in tottering mice. J. Neurophysiol. 2021, 125, 398–407. [Google Scholar] [CrossRef] [PubMed]
  34. Titley, H.K.; Watkins, G.V.; Lin, C.; Weiss, C.; McCarthy, M.; Disterhoft, J.F.; Hansel, C. Intrinsic excitability increase in cerebellar Purkinje cells following delay eyeblink conditioning in mice. J. Neurosci. 2020, 40, 2038–2046. [Google Scholar] [CrossRef]
  35. Najafi, F.; Medina, J.F. Bidirectional short-term plasticity during single-trial learning of cerebellar-driven eyelid movements in mice. Neurobiol. Learn. Mem. 2019, 170, 107097. [Google Scholar] [CrossRef] [PubMed]
  36. López-Ramos, J.C.; Houdek, Z.; Cendelín, J.; Vožeh, F.; Delgado-García, J.M. Timing correlations between cerebellar interpositus neuronal firing and classically conditioned eyelid responses in wild-type and Lurcher mice. Sci. Rep. 2018, 8, 10697. [Google Scholar] [CrossRef] [PubMed]
  37. Berger, T.W.; Alger, B.; Thompson, R.F. Neuronal substrate of classical conditioning in the hippocampus. Science 1976, 192, 483–485. [Google Scholar] [CrossRef]
  38. Solomon, P.R.; Moore, J.W. Latent inhibition and stimulus generalization of the classically conditioned nictitating membrane response in rabbits (Oryctolagus cuniculus) following dorsal hippocampal ablation. J. Comp. Physiol. Psychol. 1975, 89, 1192–1203. [Google Scholar] [CrossRef] [PubMed]
  39. McCormick, D.A.; Thompson, R.F. Neuronal responses of the rabbit cerebellum during acquisition and performance of a classically conditioned nictitating membrane-eyelid response. J. Neurosci. 1984, 4, 2811–2822. [Google Scholar] [CrossRef] [PubMed]
  40. McCormick, D.A.; Thompson, R.F. Cerebellum: Essential involvement in the classically conditioned eyelid response. Science 1984, 223, 296–299. [Google Scholar] [CrossRef] [PubMed]
  41. McCormick, D.A.; Clark, G.A.; Lavond, D.G.; Thompson, R.F. Initial localization of the memory trace for a basic form of learning. Proc. Natl. Acad. Sci. USA 1982, 79, 2731–2735. [Google Scholar] [CrossRef]
  42. Yeo, C.H.; Hardiman, M.J.; Glickstein, M. Classical conditioning of the nictitating membrane response of the rabbit. I. Lesions of the cerebellar nuclei. Exp. Brain Res. 1985, 60, 87–98. [Google Scholar] [CrossRef]
  43. Yeo, C.H.; Hardiman, M.J.; Glickstein, M. Discrete lesions of the cerebellar cortex abolish the classically conditioned nictitating membrane response of the rabbit. Behav. Brain Res. 1984, 13, 261–266. [Google Scholar] [CrossRef]
  44. Schreurs, B.G. Changes in cerebellar intrinsic neuronal excitability and synaptic plasticity result from eyeblink conditioning. Neurobiol. Learn. Mem. 2019, 166, 107094. [Google Scholar] [CrossRef] [PubMed]
  45. Baumann, O.; Borra, R.J.; Bower, J.M.; Cullen, K.E.; Habas, C.; Ivry, R.B.; Leggio, M.; Mattingley, J.B.; Molinari, M.; Moulton, E.A.; et al. Consensus paper: The role of the cerebellum in perceptual processes. Cerebellum 2015, 14, 197–220. [Google Scholar] [CrossRef] [PubMed]
  46. Freeman, J.H.; Steinmetz, A.B. Neural circuitry and plasticity mechanisms underlying delay eyeblink conditioning. Learn. Mem. 2011, 18, 666–677. [Google Scholar] [CrossRef] [PubMed]
  47. Timmann, D.; Drepper, J.; Frings, M.; Maschke, M.; Richter, S.; Gerwig, M.; Kolb, F.P. The human cerebellum contributes to motor, emotional and cognitive associative learning: A review. Cortex 2010, 46, 845–857. [Google Scholar] [CrossRef] [PubMed]
  48. Delgado-Garcia, J.M.; Gruart, A. The role of interpositus nucleus in eyelid conditioned responses. Cerebellum 2002, 1, 289–308. [Google Scholar] [CrossRef]
  49. Steinmetz, J.E. Brain substrates of classical eyeblink conditioning: A highly localized but also distributed system. Behav. Brain Res. 2000, 110, 13–24. [Google Scholar] [CrossRef]
  50. Lavond, D.G.; Kim, J.J.; Thompson, R.F. Mammalian brain substrates of aversive classical conditioning. Annu. Rev. Psychol. 1993, 44, 317–342. [Google Scholar] [CrossRef]
  51. Thurling, M.; Kahl, F.; Maderwald, S.; Stefanescu, R.M.; Schlamann, M.; Boele, H.J.; De Zeeuw, C.I.; Diedrichsen, J.; Ladd, M.E.; Koekkoek, S.K.; et al. Cerebellar cortex and cerebellar nuclei are concomitantly activated during eyeblink conditioning: A 7T fMRI study in humans. J. Neurosci. 2015, 35, 1228–1239. [Google Scholar] [CrossRef]
  52. Heiney, S.A.; Wohl, M.P.; Chettih, S.N.; Ruffolo, L.I.; Medina, J.F. Cerebellar-dependent expression of motor learning during eyeblink conditioning in head-fixed mice. J. Neurosci. 2014, 34, 14845–14853. [Google Scholar] [CrossRef] [PubMed]
  53. Boele, H.-J.; Koekkoek, S.K.E.; De Zeeuw, C.I. Cerebellar and extracerebellar involvement in mouse eyeblink conditioning; the ACDC model. Front. Cell. Neurosci. 2010, 3, 19. [Google Scholar] [CrossRef] [PubMed]
  54. Pakaprot, N.; Kim, S.; Thompson, R.F. The role of the cerebellar Interpositus nucleus in short and long term memory for trace eyeblink conditioning. Behav. Neurosci. 2009, 123, 54–61. [Google Scholar] [CrossRef]
  55. Mojtahedian, S.; Kogan, D.R.; Kanzawa, S.A.; Thompson, R.F.; Lavond, D.G. Dissociation of conditioned eye and limb responses in the cerebellar interpositus. Physiol. Behav. 2007, 91, 9–14. [Google Scholar] [CrossRef] [PubMed]
  56. Freeman, J.H., Jr.; Halverson, H.E.; Poremba, A. Differential effects of cerebellar inactivation on eyeblink conditioned excitation and inhibition. J. Neurosci. 2005, 25, 889–895. [Google Scholar] [CrossRef]
  57. Christian, K.M.; Thompson, R.F. Neural substrates of eyeblink conditioning: Acquisition and retention. Learn. Mem. 2003, 11, 427–455. [Google Scholar] [CrossRef] [PubMed]
  58. Attwell, P.J.E.; Ivarsson, M.; Millar, L.; Yeo, C.H. Cerebellar mechanisms in eyeblink conditioning. Ann. N. Y. Acad. Sci. 2002, 978, 79–92. [Google Scholar] [CrossRef]
  59. Wikgren, J.; Korhonen, T. Inactivation of the interpositus nucleus blocks the conditioned response acquired by a somatosensory conditioned stimulus in rabbit eyeblink conditioning. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 2001, 25, 1161–1168. [Google Scholar] [CrossRef] [PubMed]
  60. Welsh, J.P.; Harvey, J.A. Cerebellar lesions and the nictitating membrane reflex: Performance deficits of the conditioned and unconditioned response. J. Neurosci. 1989, 9, 299–311. [Google Scholar] [CrossRef]
  61. Delgado-Garcia, J.M.; Gruart, A. Building new motor responses: Eyelid conditioning revisited. Trends Neurosci. 2006, 29, 330–338. [Google Scholar] [CrossRef]
  62. Kelly, T.M.; Zuo, C.C.; Bloedel, J.R. Classical conditioning of the eyelid reflex in the decerebrate-decerebellate rabbit. Behav. Brain Res. 1990, 38, 7–18. [Google Scholar] [CrossRef]
  63. Perciavalle, V.; Apps, R.; Bracha, V.; Delgado-Garcia, J.M.; Gibson, A.R.; Leggio, M.; Carrel, A.J.; Cerminara, N.; Coco, M.; Gruart, A.; et al. Consensus paper: Current views on the role of cerebellar interpositus nucleus in movement control and emotion. Cerebellum 2013, 12, 738–757. [Google Scholar] [CrossRef]
  64. Attwell, P.J.E.; Rahman, S.; Yeo, C.H. Acquisition of eyeblink conditioning is critically dependent on normal function in cerebellar cortical lobule HVI. J. Neurosci. 2001, 21, 5715–5722. [Google Scholar] [CrossRef]
  65. Gruart, A.; Yeo, C.H. Cerebellar cortex and eyeblink conditioning: Bilateral regulation of conditioned responses. Exp. Brain Res. 1995, 104, 431–448. [Google Scholar] [CrossRef] [PubMed]
  66. Mostofi, A.; Holtzman, T.; Grout, A.S.; Yeo, C.H.; Edgley, S.A. Electrophysiological localization of eyeblink-related microzones in rabbit cerebellar cortex. J. Neurosci. 2010, 30, 8920–8934. [Google Scholar] [CrossRef]
  67. Nicholson, D.A.; Freeman, J.H., Jr. Developmental changes in eyeblink conditioning and simple spike activity in the cerebellar cortex. Dev. Psychobiol. 2003, 44, 45–57. [Google Scholar] [CrossRef]
  68. Miller, M.J.; Chen, N.-K.; Li, L.; Tom, B.; Weiss, C.; Disterhoft, J.F.; Wyrwicz, A.M. fMRI of the conscious rabbit during unilateral classical eyeblink conditioning reveals bilateral cerebellar activation. J. Neurosci. 2003, 23, 11753–11758. [Google Scholar] [CrossRef] [PubMed]
  69. Lavond, D.G.; Steinmetz, J.E.; Yokaitis, M.H.; Thompson, R.F. Reacquisition of classical conditioning after removal of cerebellar cortex. Exp. Brain Res. 1987, 67, 569–593. [Google Scholar] [CrossRef]
  70. Harvey, J.A.; Welsh, J.P.; Yeo, C.H.; Romano, A.G. Recoverable and nonrecoverable deficits in conditioned responses after cerebellar cortical lesions. J. Neurosci. 1993, 13, 1624–1635. [Google Scholar] [CrossRef] [PubMed]
  71. Rodriguez, F.; Duran, E.; Gomez, A.; Ocanan, F.M.; Alvarez, E.; Jimenez-Moya, F. Cognitive and emotional functions of the teleost fish cerebellum. Brain Res. Bull. 2005, 66, 365–370. [Google Scholar] [CrossRef] [PubMed]
  72. Keifer, J.; Armstrong, K.E.; Houk, J.C. In vitro classical conditioning of abducens nerve discharge in turtles. J. Neurosci. 1995, 15, 5036–5048. [Google Scholar] [CrossRef] [PubMed]
  73. Skelton, R.W. Bilateral cerebellar lesions disrupt conditioned eyelid responses in unrestrained rats. Behav. Neurosci. 1988, 102, 586–590. [Google Scholar] [CrossRef] [PubMed]
  74. Chen, L.; Bao, S.; Lockard, J.M.; Kim, J.J.; Thompson, R.F. Impaired classical eyeblink conditioning in cerebellar-lesioned and Purkinje cell degeneration (pcd) mutant mice. J. Neurosci. 1996, 16, 2829–2838. [Google Scholar] [CrossRef] [PubMed]
  75. Li, R.; Li, Q.; Chu, X.; Li, L.; Li, X.; Li, J.; Yang, Z.; Xu, M.; Luo, C.; Zhang, K. Role of cerebellar cortex in associative learning and memory in guinea pigs. Open Life Sci. 2022, 17, 1208–1216. [Google Scholar] [CrossRef] [PubMed]
  76. Hu, B.; Yang, L.; Huang, L.-S.; Chen, H.; Zeng, Y.; Feng, H.; Sui, J.-F. Effect of cerebellar reversible inactivations on the acquisition of trace conditioned eyeblink responses in guinea pigs: Comparison of short and long trace intervals. Neurosci. Lett. 2009, 459, 41–45. [Google Scholar] [CrossRef] [PubMed]
  77. Kotani, S.; Kawahara, S.; Kirino, Y. Purkinje cell activity during classical eyeblink conditioning in decerebrate guinea pigs. Brain Res. 2006, 1068, 70–81. [Google Scholar] [CrossRef] [PubMed]
  78. Delgado-Garcia, J.M.; Gruart, A. Firing activities of identified posterior interpositus nucleus neurons during associative learning in behaving cats. Brain Res. Rev. 2005, 49, 367–376. [Google Scholar] [CrossRef]
  79. Trigo, J.A.; Roa, L.; Gruart, A.; Delgado-Garcia, J.M. A kinetic study of blinking responses in cats. J. Physiol. 2003, 549, 195–205. [Google Scholar] [CrossRef]
  80. Gruart, A.; Blazquez, P.; Delgado-Garcia, J.M. Kinematics of spontaneous, reflex and conditioned eyelid movements in the alert cat. J. Neurophysiol. 1995, 74, 226–248. [Google Scholar] [CrossRef]
  81. Gerwig, M.; Dimitrova, A.; Kolb, F.P.; Maschke, M.; Brol, B.; Kunnel, A.; Boring, D.; Thilmann, A.F.; Forsting, M.; Diener, H.C.; et al. Comparison of eyeblink conditioning in patients with superior and posterior inferior cerebellar lesions. Brain 2003, 126, 71–94. [Google Scholar] [CrossRef]
  82. Gonzalez-Joekes, J. Transsynaptic Delineation of the Rabbit Eyeblink Premotor Pathway: Identifying Anterior Interpositus Neurons and Their Synaptic Changes as a Function of Learning. Unpublished Doctoral Dissertation. Ph.D. Thesis, West Virginia University, Morgantown, WV, USA, 2014. [Google Scholar]
  83. Freeman, J.H., Jr.; Nicholson, D.A. Developmental changes in eye-blink conditioning and neuronal activity in the cerebellar interpositus nucleus. J. Neurosci. 2000, 20, 813–819. [Google Scholar] [CrossRef] [PubMed]
  84. Weeks, A.C.W.; Connor, S.; Hinchcliff, R.; LeBoutillier, J.C.; Thompson, R.F.; Petit, T.L. Eye-blink conditioning is associated with changes in synaptic ultrastructure in the rabbit interpositus nuclei. Learn. Mem. 2007, 14, 385–389. [Google Scholar] [CrossRef] [PubMed]
  85. Kleim, J.A.; Freeman, J.H., Jr.; Bruneau, R.; Nolan, B.C.; Cooper, N.R.; Zook, A.; Walters, D. Synapse formation is associated with memory storage in the cerebellum. Proc. Natl. Acad. Sci. USA 2002, 99, 13228–13231. [Google Scholar] [CrossRef]
  86. Broersen, R.; Albergaria, C.; Carulli, D.; Carey, M.R.; Canto, C.B.; De Zeeuw, C.I. Synaptic mechanisms for associative learning in the cerebellar nuclei. Nat. Commun. 2023, 14, 7459. [Google Scholar] [CrossRef] [PubMed]
  87. Wang, D.; Smith-Bell, C.A.; Burhans, L.B.; O’Dell, D.E.; Bell, R.W.; Schreurs, B.G. Changes in membrane properties of rat deep cerebellar nuclear projection neurons during acquisition of eyeblink conditioning. Proc. Natl. Acad. Sci. USA 2018, 115, E9419–E9428. [Google Scholar] [CrossRef] [PubMed]
  88. Debanne, D.; Inglebert, Y.; Russier, M. Plasticity of intrinsic neuronal excitability. Curr. Opin. Neurobiol. 2019, 54, 73–82. [Google Scholar] [CrossRef] [PubMed]
  89. Brown, J.T.; Chin, J.; Leiser, S.C.; Pangalos, M.N.; Randall, A.D. Altered intrinsic neuronal excitability and reduced Na+ currents in a mouse model of Alzheimer’s disease. Neurobiol. Aging 2011, 32, 2109.e1–2109.e14. [Google Scholar] [CrossRef] [PubMed]
  90. Belmeguenai, A.; Hosy, E.; Bengtsson, F.; Pedroarena, C.M.; Piochon, C.; Teuling, E.; He, Q.; Ohtsuki, G.; De Jeu, M.T.; Elgersma, Y.; et al. Intrinsic plasticity complements long-term potentiation in parallel fiber input gain control in cerebellar Purkinje cells. J. Neurosci. 2010, 30, 13630–13643. [Google Scholar] [CrossRef] [PubMed]
  91. Goudarzi, I.; Kaffashian, M.; Shabani, M.; Haghdoost-Yazdi, H.; Behzadi, G.; Janahmadi, M. In vivo 4-aminopyridine treatment alters the neurotoxin 3-acetylpyridine-induced plastic changes in intrinsic electrophysiological properties of rat cerebellar Purkinje neurones. Eur. J. Pharmacol. 2010, 642, 56–65. [Google Scholar] [CrossRef] [PubMed]
  92. Kim, J.; Jung, S.-C.; Clemens, A.M.; Petralia, R.S.; Hoffman, D.A. Regulation of dendritic excitability by activity-dependent trafficking of the A-type K+ channel subunit Kv4.2 in hippocampal neurons. Neuron 2007, 54, 933–947. [Google Scholar] [CrossRef] [PubMed]
  93. Gittis, A.H.; du Lac, S. Intrinsic and synaptic plasticity in the vestibular system. Curr. Opin. Neurobiol. 2006, 16, 385–390. [Google Scholar] [CrossRef]
  94. Plant, L.D.; Webster, N.J.; Boyle, J.P.; Ramsden, M.; Frier, D.B.; Peers, C.; Pearson, H.A. Amyloid b peptide as a physiological modulator of neuronal “A”-type K+ current. Neurobiol. Aging 2006, 27, 1673–1683. [Google Scholar] [CrossRef]
  95. Straka, H.; Vibert, N.; Vidal, P.P.; Moore, L.E.; Dutia, M.B. Intrinsic membrane properties of vertebrate vestibular neurons: Function, development and plasticity. Prog. Neurobiol. 2005, 76, 349–392. [Google Scholar] [CrossRef] [PubMed]
  96. Cantrell, A.R.; Catterall, W.A. Neuromodulation of Na+ channels: An unexpected form of cellular plasticity. Nat. Rev. Neurosci. 2004, 2, 397–407. [Google Scholar] [CrossRef] [PubMed]
  97. Strauss, U.; Kole, M.H.P.; Bräuer, A.U.; Pahnke, J.; Bajorat, R.; Rolfs, A.; Nitsch, R.; Deisz, R.A. An impaired neocortical Ih is associated with enhanced excitability and absence epilepsy. Eur. J. Neurosci. 2004, 19, 3048–3058. [Google Scholar] [CrossRef] [PubMed]
  98. Aizenman, C.D.; Huang, E.J.; Linden, D.J. Morphological correlates of intrinsic electrical excitability in neurons of the deep cerebellar nuclei. J. Neurophysiol. 2003, 89, 1738–1747. [Google Scholar] [CrossRef] [PubMed]
  99. Johnston, D.; Christie, B.R.; Frick, A.; Gray, R.; Hoffman, D.A.; Schexnayder, L.K.; Watanabe, S.; Yuan, L.-L. Active dendrites, potassium channels and synaptic plasticity. Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 2003, 358, 667–674. [Google Scholar] [CrossRef]
  100. Nelson, A.B.; Krispel, C.M.; Sekirnjak, C.; du Lac, S. Long-lasting increases in intrinsic excitability triggered by inhibition. Neuron 2003, 40, 609–620. [Google Scholar] [CrossRef] [PubMed]
  101. Daoudal, G.; Debanne, D. Long-term plasticity of intrinsic excitability: Learning rules and mechanisms. Learn. Mem. 2003, 10, 456–465. [Google Scholar] [CrossRef]
  102. Schrader, L.A.; Anderson, A.E.; Varga, A.W.; Levy, M.; Sweatt, J.D. The other half of Hebb: K+ channels and the regulation of neuronal excitability in the hippocampus. Mol. Neurobiol. 2002, 25, 51–66. [Google Scholar]
  103. Bekkers, J.M.; Delaney, A.J. Modulation of excitability by a-dendrotoxin-sensitive potassium channels in neocortical pyramidal neurons. J. Neurosci. 2001, 21, 6553–6560. [Google Scholar] [CrossRef]
  104. Armano, S.; Rossi, P.; D’Angelo, E. Long-term potentiation of intrinsic excitability at the mossy fiber-granule cell synapse of rat cerebellum. J. Neurosci. 2000, 20, 5208–5216. [Google Scholar] [CrossRef]
  105. Pongs, O. Voltage-gated potassium channels: From hyperexcitability to excitement. FEBS Lett. 1999, 452, 31–35. [Google Scholar] [CrossRef]
  106. Armstrong, C.M.; Hille, B. Voltage-gated ion channels and electrical excitability. Neuron 1998, 20, 371–380. [Google Scholar] [CrossRef]
  107. Gruol, D.L.; Dionne, V.E.; Yool, A.J. Multiple voltage-sensitive K+ channels regulate dendritic excitability in cerebellar Purkinje neurons. Neurosci. Lett. 1989, 97, 97–102. [Google Scholar] [CrossRef]
  108. Adams, P.R.; Galvan, M. Voltage-dependent currents of vertebrate neurons and their role in membrane excitability. In Advances in Neurology, 44th ed.; Delgado-Escueta, A.V., Ed.; Raven Press: New York, NY, USA, 1986; pp. 137–170. [Google Scholar]
  109. dos-Santos, R.C.; Sweeten, B.L.; Stelly, C.E.; Tasker, J.G. The Neuroendocrine Impact of Acute Stress on Synaptic Plasticity. Endocrinology 2023, 164, bqad149. [Google Scholar] [CrossRef]
  110. Huang, H.; Shakkottai, V.G. Targeting ion channels and purkinje neuron intrinsic membrane excitability as a therapeutic strategy for cerebellar ataxia. Life 2023, 13, 1350. [Google Scholar] [CrossRef]
  111. Carzoli, K.L.; Kogias, G.; Fawcett-Patel, J.; Liu, S.J. Cerebellar interneurons control fear memory consolidation via learning-induced HCN plasticity. Cell Rep. 2023, 42, 113057. [Google Scholar] [CrossRef]
  112. Daou, A.; Margoliash, D. Intrinsic plasticity and birdsong learning. Neurobiol. Learn. Mem. 2021, 180, 107407. [Google Scholar] [CrossRef]
  113. Hansel, C.; Disterhoft, J.F. Why is synaptic plasticity not enough? Neurobiol. Learn. Mem. 2020, 176, 107336. [Google Scholar] [CrossRef] [PubMed]
  114. Titley, H.K.; Brunel, N.; Hansel, C. Toward a neurocentric view of learning. Neuron 2017, 95, 19–32. [Google Scholar] [CrossRef] [PubMed]
  115. Llinas, R.R. Intrinsic electrical properties of mammalian neurons and CNS function: A historical perspective. Front. Cell. Neurosci. 2014, 8, 320. [Google Scholar] [CrossRef] [PubMed]
  116. Wijesinghe, R.; Camp, A.J. Intrinsic neuronal excitability: Implications for health and disease. Biomol. Concepts 2011, 2, 247–259. [Google Scholar] [CrossRef] [PubMed]
  117. Zhang, W.; Linden, D.J. The other side of the engram: Experience-driven changes in neuronal intrinsic excitability. Nat. Rev. Neurosci. 2003, 4, 885–900. [Google Scholar] [CrossRef] [PubMed]
  118. Jang, D.C.; Chung, G.; Kim, S.K.; Kim, S.J. Dynamic alteration of intrinsic properties of the cerebellar Purkinje cell during the motor memory consolidation. Mol. Brain 2023, 16, 58. [Google Scholar] [CrossRef] [PubMed]
  119. Viet, N.M.; Wang, T.; Tran-Anh, K.; Sugihara, I. Heterogeneity of intrinsic plasticity in cerebellar Purkinje cells linked with cortical molecular zones. iScience 2022, 25, 103705. [Google Scholar] [CrossRef]
  120. Johansson, F. Intrinsic memory of temporal intervals in cerebellar Purkinje cells. Neurobiol. Learn. Mem. 2019, 166, 107103. [Google Scholar] [CrossRef] [PubMed]
  121. Jang, D.C.; Kim, S.J. Plasticity leading to cerebellum-dependent learning: Two different regions, two different types. Pflüg. Arch.-Eur. J. Physiol. 2019, 471, 927–934. [Google Scholar] [CrossRef]
  122. Shim, H.G.; Lee, Y.S.; Kim, S.J. The emerging concept of intrinsic plasticity: Activity-dependent modulation of intrinsic excitability in cerebellar purkinje cells and motor learning. Exp. Neurobiol. 2018, 27, 139–154. [Google Scholar] [CrossRef]
  123. Yang, Z.; Santamaria, F. Purkinje cell intrinsic excitability increases after synaptic long term depression. J. Neurophysiol. 2016, 116, 1208–1217. [Google Scholar] [CrossRef]
  124. Kim, S.J.; Linden, D.J. Ubiquitous plasticity and memory storage. Neuron 2007, 56, 582–592. [Google Scholar] [CrossRef]
  125. Zhu, L.; Scelfo, B.; Tempia, F.; Sacchetti, B.; Strata, P. Membrane excitability and fear conditioning in cerebellar Purkinje cell. Neuroscience 2006, 140, 801–810. [Google Scholar] [CrossRef]
  126. Grasselli, G.; Boele, H.J.; Titley, H.K.; Bradford, N.; van Beers, L.; Jay, L.; Beekhof, G.C.; Busch, S.E.; De Zeeuw, C.I.; Schonewille, M.; et al. SK2 channels in cerebellar Purkinje cells contribute to excitability modulation in motor-learning-specific memory traces. PLoS Biol. 2020, 18, e3000596. [Google Scholar] [CrossRef]
  127. Schreurs, B.G.; Gusev, P.A.; Tomsic, D.; Alkon, D.L.; Shi, T. Intracellular correlates of acquisition and long-term memory of classical conditioning in Purkinje cell dendrites in slices of rabbit cerebellar lobule HVI. J. Neurosci. 1998, 18, 5498–5507. [Google Scholar] [CrossRef]
  128. Schreurs, B.G.; Tomsic, D.; Gusev, P.A.; Alkon, D.L. Dendritic excitability microzones and occluded long-term depression after classical conditioning of the rabbit’s nictitating membrane response. J. Neurophysiol. 1997, 77, 86–92. [Google Scholar] [CrossRef]
  129. Schreurs, B.G.; Sanchez-Andres, J.V.; Alkon, D.L. Learning-specific differences in Purkinje-cell dendrites of lobule HVI (lobulus simplex): Intracellular recording in a rabbit cerebellar slice. Brain Res. 1991, 548, 18–22. [Google Scholar] [CrossRef] [PubMed]
  130. Wang, D.; Schreurs, B.G. Characteristics of IA currents in adult rabbit cerebellar Purkinje cells. Brain Res. 2006, 1096, 85–96. [Google Scholar] [CrossRef] [PubMed]
  131. Disterhoft, J.F.; Coulter, D.A.; Alkon, D.L. Conditioning-specific membrane changes of rabbit hippocampal neurons measured in vitro. Proc. Natl. Acad. Sci. USA 1986, 83, 2733–2737. [Google Scholar] [CrossRef]
  132. Oh, M.M.; Disterhoft, J.F. Increased excitability of both principal neurons and interneurons during associative learning. Neuroscientist 2015, 21, 372–384. [Google Scholar] [CrossRef]
  133. Hirano, T. Regulation and interaction of multiple types of synaptic plasticity in a Purkinje neuron and their contribution to motor learning. Cerebellum 2018, 17, 756–765. [Google Scholar] [CrossRef] [PubMed]
  134. Shim, H.G.; Jang, D.C.; Lee, J.; Chung, G.; Lee, S.; Kim, Y.G.; Jeon, D.E.; Kim, S.J. Long-term depression of intrinsic excitability accompanied by synaptic depression in cerebellar Purkinje cells. J. Neurosci. 2017, 37, 5659–5669. [Google Scholar] [CrossRef] [PubMed]
  135. Hoxha, E.; Tempia, F.; Lippiello, P.; Miniaci, M.C. Modulation, plasticity and pathophysiology of the parallel fiber-Purkinje cell synapse. Front. Synaptic Neurosci. 2016, 8, 35. [Google Scholar] [CrossRef] [PubMed]
  136. Jorntell, H. Cerebellar synaptic plasticity and the credit assignment problem. Cerebellum 2016, 15, 104–111. [Google Scholar] [CrossRef] [PubMed]
  137. D’Angelo, E.; Mapelli, L.; Casellato, C.; Garrido, J.A.; Luque, N.; Monaco, J.; Prestori, F.; Pedrocchi, A.; Ros, E. Distributed circuit plasticity: New clues for the cerebellar mechanisms of learning. Cerebellum 2015, 15, 139–151. [Google Scholar] [CrossRef] [PubMed]
  138. Hesslow, G.; Jirenhed, D.-A.; Rasmusson, A.; Johansson, F. Classical conditioning of motor responses: What is the learning mechanism? Neural Netw. 2013, 47, 81–87. [Google Scholar] [CrossRef] [PubMed]
  139. Ito, M.; Yamaguchi, K.; Nagao, S.; Yamzaki, T. Long-term depression as a model of cerebellar plasticity. Prog. Brain Res. 2014, 210, 1–30. [Google Scholar]
  140. Ohtsuki, G.; Shishikura, M.; Ozaki, A. Synergistic excitability plasticity in cerebellar functioning. FEBS J. 2020, 287, 4557–4593. [Google Scholar] [CrossRef] [PubMed]
  141. Pugh, J.R.; Raman, I.M. Nothing can be coincidence; synaptic inhibition and plasticity in the cerebellar nuclei. Trends Neurosci. 2009, 32, 170–177. [Google Scholar] [CrossRef]
  142. Pedroarena, C.M.; Schwarz, C. Efficacy and short-term plasticity at GABAergic synapses between Purkinje and cerebellar nuclei neurons. J. Neurophysiol. 2003, 89, 704–715. [Google Scholar] [CrossRef]
  143. Aizenman, C.D.; Linden, D.J. Rapid, synaptically driven increases in the excitability of cerebellar deep nuclear neurons. Nat. Neurosci. 2000, 3, 109–111. [Google Scholar] [CrossRef]
  144. Morishita, W.; Sastry, B.R. Postsynaptic mechanisms underlying long-term depression of GABAergic transmission in neurons of the deep cerebellar nuclei. J. Neurophysiol. 1996, 76, 59–68. [Google Scholar] [CrossRef]
  145. Najac, M.; Raman, I.M. Integration of Purkinje cell inhibition by cerebellar nucleo-olivary neurons. J. Neurosci. 2015, 35, 544–549. [Google Scholar] [CrossRef]
  146. Zheng, N.; Raman, I.M. Synaptic inhibition, excitation, and plasticity in neurons of the cerebellar nuclei. Cerebellum 2010, 9, 56–66. [Google Scholar] [CrossRef]
  147. Pugh, J.R.; Raman, I.M. Potentiation of mossy fiber EPSCs in the cerebellar nuclei by NMDA receptor activation followed by postinhibitory rebound current. Neuron 2006, 51, 113–123. [Google Scholar] [CrossRef]
  148. Telgkamp, P.; Raman, I.M. Depression of inhibitory synaptic transmission between Purkinje cells and neurons of the cerebellar nuclei. J. Neurosci. 2002, 22, 8447–8457. [Google Scholar] [CrossRef] [PubMed]
  149. Raman, I.M.; Gustafson, A.E.; Padgett, D. Ionic currents and spontaneous firing in neurons isolated from the cerebellar nuclei. J. Neurosci. 2000, 20, 9004–9016. [Google Scholar] [CrossRef]
  150. Marr, D. A theory of cerebellar cortex. J. Physiol. 1969, 202, 437–470. [Google Scholar] [CrossRef] [PubMed]
  151. Albus, J.S. A theory of cerebellar function. Math. Biosci. 1971, 10, 25–61. [Google Scholar] [CrossRef]
  152. Ito, M.; Sakurai, M.; Tongroach, P. Climbing fibre induced depression of both mossy fibre responsiveness and glutamate sensitivity of cerebellar Purkinje cells. J. Physiol. 1982, 324, 113–134. [Google Scholar] [CrossRef]
  153. Ito, M.; Kano, M. Long-lasting depression of parallel fiber-Purkinje cell transmission induced by conjunctive stimulation of parallel fibers and climbing fibers in the cerebellar cortex. Neurosci. Lett. 1982, 33, 253–258. [Google Scholar] [CrossRef]
  154. Gilbert, P.F.C.; Thach, W.T. Purkinje cell activity during motor learning. Brain Res. 1977, 128, 309–328. [Google Scholar] [CrossRef]
  155. Pugh, J.R.; Raman, I.M. Mechanisms of potentiation of mossy fiber EPSCs in the cerebellar nuclei by coincident synaptic excitation and inhibition. J. Neurosci. 2008, 28, 10549–10560. [Google Scholar] [CrossRef]
  156. Aizenman, C.D.; Huang, E.J.; Manis, P.B.; Linden, D.J. Use-Dependent Changes in Synaptic Strength at the Purkinje Cell to Deep Nuclear Synapse; Gerrits, N.M., Ruigrok, T.J.H., De Zeeuw, C.I., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2000; Volume 124, pp. 257–273. [Google Scholar]
  157. Aizenman, C.D.; Manis, P.B.; Linden, D.J. Polarity of long-term synaptic gain change is related to postsynaptic spike firing at a cerebellar inhibitory synapse. Neuron 1998, 21, 827–835. [Google Scholar] [CrossRef]
  158. Ouardouz, M.; Sastry, B.R. Mechanisms underlying LTP of inhibitory synaptic transmission in the deep cerebellar nuclei. J. Neurophysiol. 2000, 84, 1414–1421. [Google Scholar] [CrossRef]
  159. Ito, M. Cerebellar learning in the vestibulo-ocular reflex. Trends Cogn. Sci. 1998, 2, 313–321. [Google Scholar] [CrossRef]
  160. Takehara-Nishiuchi, K. The anatomy and physiology of eyeblink classical conditioning. Curr. Top. Behav. Neurosci. 2018, 37, 297–323. [Google Scholar] [CrossRef]
  161. Johansson, F.; Jirenhed, D.A.; Rasmussen, A.; Zucca, R.; Hesslow, G. Absence of parallel fibre to purkinje cell LTD during eyeblink conditioning. Sci. Rep. 2018, 8, 14777. [Google Scholar] [CrossRef]
  162. Thompson, R.F.; Steinmetz, J.E. The role of the cerebellum in classical conditioning of discrete behavioral responses. Neuroscience 2009, 162, 732–755. [Google Scholar] [CrossRef] [PubMed]
  163. Koekkoek, S.K.E.; Hulscher, H.C.; Dortland, B.R.; Hensbroek, R.A.; Elgersma, Y.; Ruigrok, T.J.H.; De Zeeuw, C.I. Cerebellar LTD and learning-dependent timing of conditioned eyelid responses. Science 2003, 301, 1736–1739. [Google Scholar] [CrossRef] [PubMed]
  164. Ito, M. Cerebellar long-term depression: Characterization, signal transduction, and functional roles. Physiol. Rev. 2001, 81, 1143–1195. [Google Scholar] [CrossRef] [PubMed]
  165. Mauk, M.D.; Garcia, K.S.; Medina, J.F.; Steele, P.M. Does cerebellar LTD mediate motor learning? Toward a resolution without a smoking gun. Neuron 1998, 20, 359–362. [Google Scholar] [CrossRef] [PubMed]
  166. Cajal, S.R. The Croonian lecture—La fine structure des centres nerveux. Proc. R. Soc. Lond. 1894, 55, 444–468. [Google Scholar]
  167. Yuste, R.; Tank, D.W. Dendritic integrations in mammalian neurons, a century after Cajal. Neuron 1996, 16, 701–716. [Google Scholar] [CrossRef] [PubMed]
  168. Sotelo, C. Viewing the cerebellum through the eyes of Ramon Y Cajal. Cerebellum 2008, 7, 517–522. [Google Scholar] [CrossRef] [PubMed]
  169. Heysieattalab, S.; Lee, K.H.; Liu, Y.; Wang, Y.; Foy, M.R.; Bi, X.; Baudry, M. Impaired cerebellar plasticity and eye-blink conditioning in calpain-1 knock-out mice. Neurobiol. Learn. Mem. 2020, 170, 106995. [Google Scholar] [CrossRef] [PubMed]
  170. Connor, S.; LeBoutillier, J.C.; Petit, T.L.; Bloomfield, J.; Thompson, R.F.; Weeks, A.C.W. Eyeblink conditioning leads to fewer synapses in the rabbit cerebellar cortex. Behav. Neurosci. 2009, 123, 856–862. [Google Scholar] [CrossRef]
  171. Miyata, M.; Kim, H.-T.; Hashimoto, K.; Lee, T.-K.; Cho, S.-Y.; Jiang, H.; Wu, Y.; Jun, K.; Wu, D.; Kano, M.; et al. Deficient long-term synaptic depression in the rostral cerebellum correlated with impaired motor learning in phospholipase C b4 mutant mice. Eur. J. Neurosci. 2001, 13, 1945–1954. [Google Scholar] [CrossRef] [PubMed]
  172. Schonewille, M.; Belmeguenai, A.; Koekkoek, S.K.; Houtman, S.H.; Boele, H.J.; van Beugen, B.J.; Gao, Z.; Badura, A.; Ohtsuki, G.; Amerika, W.E.; et al. Purkinje cell-specific knockout of the protein phosphatase PP2B impairs potentiation and cerebellar motor learning. Neuron 2010, 67, 618–628. [Google Scholar] [CrossRef] [PubMed]
  173. Welsh, J.P.; Yamaguchi, H.; Zeng, X.-H.; Kojo, M.; Nakada, Y.; Takagi, A.; Sugimori, M.; Llinas, R.R. Normal motor learning during pharmacological prevention of Purkinje cell long-term depression. Proc. Natl. Acad. Sci. USA 2006, 102, 17166–17171. [Google Scholar] [CrossRef]
  174. Carey, M.R.; Lisberger, S.G. Embarrassed, but not depressed: Eye opening lessons for cerebellar learning. Neuron 2002, 35, 223–226. [Google Scholar] [CrossRef] [PubMed]
  175. Schonewille, M.; Gao, Z.; Boele, H.J.; Vinueza Veloz, M.F.; Amerika, W.E.; Simek, A.A.; De Jeu, M.T.; Steinberg, J.P.; Takamiya, K.; Hoebeek, F.E.; et al. Reevaluating the Role of LTD in Cerebellar Motor Learning. Neuron 2011, 70, 43–50. [Google Scholar] [CrossRef]
  176. Tseng, W.; Guan, R.; Disterhoft, J.F.; Weiss, C. Trace eyeblink conditioning is hippocampally dependent in mice. Hippocampus 2004, 14, 58–65. [Google Scholar] [CrossRef]
  177. McEchron, M.D.; Tseng, W.; Disterhoft, J.F. Single neurons in CA1 hippocampus encode trace interval duration during trace heart rate (fear) conditioning in rabbit. J. Neurosci. 2003, 23, 1535–1547. [Google Scholar] [CrossRef] [PubMed]
  178. Clark, R.E.; Zola, S. Trace eyeblink classical conditioning in the monkey: A nonsurgical method and behavioral analysis. Behav. Neurosci. 1998, 112, 1062–1068. [Google Scholar] [CrossRef] [PubMed]
  179. McEchron, M.D.; Bouwmeester, H.; Tseng, W.; Weiss, C.; Disterhoft, J.F. Hippocampectomy disrupts auditory trace fear conditioning and contextual fear conditioning in the rat. Hippocampus 1998, 8, 638–646. [Google Scholar] [CrossRef]
  180. McGlinchey-Berroth, R.; Carillo, M.C.; Gabrieli, J.D.E.; Brawn, C.M.; Disterhoft, J.F. Impaired trace eyeblink conditioning in bilateral, medial-temporal lobe amnesia. Behav. Neurosci. 1997, 111, 873–882. [Google Scholar] [CrossRef] [PubMed]
  181. Moye, T.B.; Rudy, J.W. Ontogenesis of trace conditioning in young rats: Dissociation of associative and memory processes. Dev. Psychobiol. 1987, 20, 405–414. [Google Scholar] [CrossRef]
  182. Solomon, P.R.; Vander Schaaf, E.R.; Thompson, R.F.; Weisz, D.J. Hippocampus and trace conditioning of the rabbit’s classically conditioned nictitating membrane response. Behav. Neurosci. 1986, 100, 729–744. [Google Scholar] [CrossRef] [PubMed]
  183. Schneiderman, N. Interstimulus interval function of the nictitating membrane response of the rabbit under delay versus trace conditioning. J. Comp. Physiol. Psychol. 1966, 62, 397–402. [Google Scholar] [CrossRef]
  184. Thompson, R.F. The neurobiology of learning and memory. Science 1986, 233, 941–947. [Google Scholar] [CrossRef] [PubMed]
  185. Huang, C.M.; Liu, G. Electrophysiological mapping of the auditory areas in the cerebellum of the cat. Brain Res. 1985, 335, 121–129. [Google Scholar] [CrossRef]
  186. Gould, T.J.; Sears, L.L.; Steinmetz, J.E. Possible CS and US pathways for rabbit classical eyelid conditioning: Electrophysiological evidence for projections from the pontine nuclei and inferior olive to cerebellar cortex and nuclei. Behav. Neural Biol. 1993, 60, 172–185. [Google Scholar] [CrossRef]
  187. Siegel, J.J.; Mauk, M.D. Persistent activity in prefrontal cortex during trace eyelid conditioning: Dissociating responses that reflect cerebellar output from those that do not. J. Neurosci. 2013, 33, 15272–15284. [Google Scholar] [CrossRef]
  188. Hattori, S.; Yoon, T.; Disterhoft, J.F.; Weiss, C. Functional reorganization of a prefrontal cortical network mediating consolidation of trace eyeblink conditioning. J. Neurosci. 2014, 34, 1432–1445. [Google Scholar] [CrossRef]
  189. Yu, X.T.; Yu, J.; Choi, A.; Takehara-Nishiuchi, K. Lateral entorhinal cortex supports the development of prefrontal network activity that bridges temporally discontiguous stimuli. Hippocampus 2021, 31, 1285–1299. [Google Scholar] [CrossRef]
  190. Kehoe, E.J.; Ludvig, E.A.; Sutton, R.S. Timing and cue competition in conditioning of the nictitating membrane response of the rabbit (Oryctolagus cuniculus). Learn. Mem. 2013, 20, 97–102. [Google Scholar] [CrossRef]
  191. Gormezano, I.; Kehoe, E.J. Associative transfer in classical conditioning to serial compounds. In Quantitative Analyses of Behavior; Commons, M.L., Herrnstein, R.J., Wagner, A.R., Eds.; Ballinger: Cambridge, MA, USA, 1984; pp. 297–322. [Google Scholar]
  192. Kehoe, E.J.; Feyer, A.M.; Moses, J.L. Second-order conditioning of the rabbit’s nictitating membrane response as a function of the CS2-CS1 and CS1-US intervals. Anim. Learn. Behav. 1981, 9, 304–314. [Google Scholar] [CrossRef]
  193. Kehoe, E.J.; Gibbs, C.M.; Garcia, E.; Gormezano, I. Associative transfer and stimulus selection in classical conditioning of the rabbit’s nictitating membrane response to serial compound CSs. J. Exp. Psychol. Anim. Behav. Process. 1979, 5, 1–18. [Google Scholar] [CrossRef]
  194. Kehoe, E.J.; Graham-Clarke, P.; Schreurs, B.G. Temporal patterns of the rabbit’s nictitating membrane response to compound and component stimuli under mixed CS-US intervals. Behav. Neurosci. 1989, 103, 283–295. [Google Scholar] [CrossRef]
  195. Schreurs, B.G.; Alkon, D.L. US-US conditioning of the rabbit’s nictitating membrane response: Emergence of a conditioned response without alpha conditioning. Psychobiology 1990, 18, 312–320. [Google Scholar] [CrossRef]
  196. Schreurs, B.G.; Burhans, L.B.; Smith-Bell, C.A.; Mrowka, S.W.; Wang, D. Ontogeny of trace eyeblink conditioning to shock-shock pairings in the rat pup. Behav. Neurosci. 2013, 127, 114–120. [Google Scholar] [CrossRef] [PubMed]
  197. Schmolesky, M.T.; De Zeeuw, C.I.; Hansel, C. Climbing fiber synaptic plasticity and modifications in Purkinje cell excitability. Prog. Brain Res. 2005, 148, 81–94. [Google Scholar] [PubMed]
  198. Hansel, C.; Linden, D.J. Long-term depression of the cerebellar climbing fiber-Purkinje neuron synapse. Neuron 2000, 26, 473–482. [Google Scholar] [CrossRef] [PubMed]
  199. Mozzachiodi, R.; Byrne, J.H. More than synaptic plasticity: Role of nonsynaptic plasticity in learning and memory. Trends Neurosci. 2010, 33, 17–26. [Google Scholar] [CrossRef] [PubMed]
  200. Hirono, M.; Karube, F.; Yanagawa, Y. Modulatory effects of monoamines and perineuronal nets on output of cerebellar Purkinje cells. Front. Neural Circuits 2021, 15, 661899. [Google Scholar] [CrossRef] [PubMed]
  201. O’Dell, D.E.; Schreurs, B.G.; Smith-Bell, C.; Wang, D. Disruption of rat deep cerebellar perineuronal net alters eyeblink conditioning and neuronal electrophysiology. Neurobiol. Learn. Mem. 2021, 177, 107358. [Google Scholar] [CrossRef]
  202. Carulli, D.; Broersen, R.; de Winter, F.; Muir, E.M.; Meskovic, M.; de Waal, M.; de Vries, S.; Boele, H.J.; Canto, C.B.; De Zeeuw, C.I.; et al. Cerebellar plasticity and associative memories are controlled by perineuronal nets. Proc. Natl. Acad. Sci. USA 2020, 117, 6855–6865. [Google Scholar] [CrossRef]
  203. Hirono, M.; Watanabe, S.; Karube, F.; Fujiyama, F.; Kawahara, S.; Nagao, S.; Yanagawa, Y.; Misonou, H. Perineuronal nets in the deep cerebellar nuclei regulate GABAergic transmission and delay eyeblink conditioning. J. Neurosci. 2018, 38, 6130–6144. [Google Scholar] [CrossRef]
  204. Celio, M.R.; Blümcke, I. Perineuronal nets--a specialized form of extracellular matrix in the adult nervous system. Brain Res. Rev. 1994, 19, 128–145. [Google Scholar] [CrossRef] [PubMed]
  205. Celio, M.R.; Spreafico, R.; De Biasi, S.; Vitellaro-Zuccarello, L. Perineuronal nets: Past and present. Trends Neurosci. 1998, 21, 510–515. [Google Scholar] [CrossRef] [PubMed]
  206. Sotelo, C. The History of the Synapse. Anat. Rec. 2020, 303, 1252–1279. [Google Scholar] [CrossRef] [PubMed]
  207. Fellin, T. Communication between neurons and astrocytes: Relevance to the modulation of synaptic and network activity. J. Neurochem. 2009, 108, 533–544. [Google Scholar] [CrossRef]
  208. Halassa, M.M.; Fellin, T.; Haydon, P.G. Tripartite synapses: Roles for astrocytic purines in the control of synaptic physiology and behavior. Neuropharmacology 2009, 57, 343–346. [Google Scholar] [CrossRef]
  209. Perea, G.; Navarrete, M.; Araque, A. Tripartite synapses: Astrocytes process and control synaptic information. Trends Neurosci. 2009, 32, 421–431. [Google Scholar] [CrossRef]
  210. Sorg, B.A.; Berretta, S.; Blacktop, J.M.; Fawcett, J.W.; Kitagawa, H.; Kwok, J.C.F.; Miquel, M. Casting a wide net: Role of perineuronal nets in neural plasticity. J. Neurosci. 2016, 36, 11459–11468. [Google Scholar] [CrossRef]
  211. Tewari, B.P.; Chaunsali, L.; Prim, C.E.; Sontheimer, H. A glial perspective on the extracellular matrix and perineuronal net remodeling in the central nervous system. Front. Cell. Neurosci. 2022, 16, 1022754. [Google Scholar] [CrossRef]
  212. Crapser, J.D.; Arreola, M.A.; Tsourmas, K.I.; Green, K.N. Microglia as hackers of the matrix: Sculpting synapses and the extracellular space. Cell. Mol. Immunol. 2021, 18, 2472–2488. [Google Scholar] [CrossRef]
  213. Ferrer-Ferrer, M.; Dityatev, A. Shaping synapses by the neural extracellular matrix. Front. Neuroanat. 2018, 12, 40. [Google Scholar] [CrossRef]
  214. Fawcett, J.W.; Kwok, J.C.F. Proteoglycan sulphation in the function of the mature central nervous system. Front. Integr. Neurosci. 2022, 16, 895493. [Google Scholar] [CrossRef] [PubMed]
  215. Carceller, H.; Gramuntell, Y.; Klimczak, P.; Nacher, J. Perineuronal nets: Subtle structures with large implications. Neuroscientist 2022, 29, 569–590. [Google Scholar] [CrossRef] [PubMed]
  216. Carulli, D.; Verhaagen, J. An extracellular perspective on CNS maturation: Perineuronal nets and the control of plasticity. Int. J. Mol. Sci. 2021, 22, 2434. [Google Scholar] [CrossRef] [PubMed]
  217. Testa, D.; Prochiantz, A.; Di Nardo, A.A. Perineuronal nets in brain physiology and disease. Semin. Cell Dev. Biol. 2019, 89, 125–135. [Google Scholar] [CrossRef] [PubMed]
  218. Song, I.; Dityatev, A. Crosstalk between glia, extracellular matrix and neurons. Brain Res. Bull. 2018, 136, 101–108. [Google Scholar] [CrossRef]
  219. Van’t Spijker, H.M.; Kwok, J.C.F. A sweet talk: The molecular systems of perineuronal nets in controlling neuronal communication. Front. Integr. Neurosci. 2017, 11, 33. [Google Scholar] [CrossRef] [PubMed]
  220. Levy, A.D.; Omar, M.H.; Koleske, A.J. Extracellular matrix control of dendritic spine and synapse structure and plasticity in adulthood. Front. Neuroanat. 2014, 8, 116. [Google Scholar] [CrossRef] [PubMed]
  221. Thach, W.T. Discharge of Purkinje and cerebellar nuclear neurons during rapidly alternating arm movements in the monkey. J. Neurophysiol. 1968, 31, 785–796. [Google Scholar] [CrossRef] [PubMed]
  222. Khan, M.M.; Wu, S.; Chen, C.H.; Regehr, W.G. Unusually slow spike frequency adaptation in deep cerebellar nuclei neurons preserves linear transformations on the subsecond timescale. J. Neurosci. 2022, 42, 7581–7593. [Google Scholar] [CrossRef]
  223. Murakami, M.; Kosaka, M.; Sato, H.; Ohtsuka, A.; Taguchi, T. The intensely positively charged perineuronal net in the adult rat brain, with special reference to its reactions to oxine, chondroitinase ABC, hyaluronidase and collagenase. Arch. Histol. Cytol. 2001, 64, 313–318. [Google Scholar] [CrossRef] [PubMed]
  224. Mueller, A.L.; Davis, A.; Sovich, S.; Carlson, S.S.; Robinson, F.R. Distribution of N-acetylgalactosamine-positive perineuronal nets in the macaque brain: Anatomy and implications. Neural Plast. 2016, 2016, 6021428. [Google Scholar] [CrossRef] [PubMed]
  225. Starkey, J.; Horstick, E.J.; Ackerman, S.D. Glial regulation of critical period plasticity. Front. Cell. Neurosci. 2023, 17, 1247335. [Google Scholar] [CrossRef]
  226. Fawcett, J.W.; Oohashi, T.; Pizzorusso, T. The roles of perineuronal nets and the perinodal extracellular matrix in neuronal function. Nat. Rev. Neurosci. 2019, 20, 451–465. [Google Scholar] [CrossRef]
  227. Foscarin, S.; Raha-Chowdhury, R.; Fawcett, J.W.; Kwok, J.C.F. Brain ageing changes proteoglycan sulfation, rendering perineuronal nets more inhibitory. Aging 2017, 9, 1607–1622. [Google Scholar] [CrossRef]
  228. Ueno, H.; Suemitsu, S.; Okamoto, M.; Matsumoto, Y.; Ishihara, T. Sensory experience-dependent formation of perineuronal nets and expression of Cat-315 immunoreactive components in the mouse somatosensory cortex. Neuroscience 2017, 355, 161–174. [Google Scholar] [CrossRef] [PubMed]
  229. Nabel, E.; Morishita, H. Regulating critical period plasticity: Insight from the visual system to fear circuitry for therapeutic interventions. Front. Psychiatry 2013, 4, 146. [Google Scholar] [CrossRef] [PubMed]
  230. Carulli, D.; Rhodes, K.E.; Fawcett, J.W. Upregulation of aggrecan, link protein 1, and hyaluronan synthases during formation of perineuronal nets in the rat cerebellum. J. Comp. Neurol. 2007, 501, 83–94. [Google Scholar] [CrossRef] [PubMed]
  231. Köppe, G.; Brückner, G.; Brauer, K.; Härtig, W.; Bigl, V. Developmental patterns of proteoglycan-containing extracellular matrix in perineuronal nets and neuropil of the postnatal rat brain. Cell Tissue Res. 1997, 288, 33–41. [Google Scholar] [CrossRef] [PubMed]
  232. Ye, Q.; Miao, Q.L. Experience-dependent development of perineuronal nets and chondroitin sulfate proteoglycan receptors in mouse visual cortex. Matrix Biol. 2013, 32, 352–363. [Google Scholar] [CrossRef]
  233. Wang, D.; Schreurs, B.G. Maturation of membrane properties of neurons in the rat deep cerebellar nuclei. Dev. Neurobiol. 2014, 74, 1268–1276. [Google Scholar] [CrossRef] [PubMed]
  234. Ohtsuki, G.; Piochon, C.; Hansel, C. Climbing fiber signaling and cerebellar gain control. Front. Cell. Neurosci. 2009, 3, 548. [Google Scholar] [CrossRef] [PubMed]
  235. Goldsberry, M.E.; Freeman, J.H. Sensory system development influences the ontogeny of trace eyeblink conditioning. Dev. Psychobiol. 2017, 59, 70–76. [Google Scholar] [CrossRef]
  236. Claflin, D.I.; Garrett, T.; Buffington, M.L. A developmental comparison of trace and delay eyeblink conditioning in rats using matching interstimulus intervals. Dev. Psychobiol. 2005, 47, 77–88. [Google Scholar] [CrossRef]
  237. Freeman, J.H., Jr.; Nicholson, D.A.; Muckler, A.S.; Rabinak, C.A.; DiPietro, N.T. Ontogeny of eyeblink conditioned response timing in rats. Behav. Neurosci. 2003, 117, 283–291. [Google Scholar] [CrossRef]
  238. Ivkovich, D.; Paczkowski, C.; Stanton, M.E. Ontogeny of delay versus trace eyeblink conditioning in the rat. Dev. Psychobiol. 2000, 36, 148–160. [Google Scholar] [CrossRef]
  239. Stanton, M.E.; Fox, G.D.; Carter, C.S. Ontogeny of the conditioned eyeblink response in rats: Acquisition or expression. Neuropharmacology 1998, 37, 623–632. [Google Scholar] [CrossRef] [PubMed]
  240. Freeman, J.H., Jr.; Rabinak, C.A.; Campolattaro, M.M. Pontine stimulation overcomes developmental limitations in the neural mechanisms of eyeblink conditioning. Learn. Mem. 2005, 12, 255–259. [Google Scholar] [CrossRef]
  241. Wingert, J.C.; Sorg, B.A. Impact of perineuronal nets on electrophysiology of parvalbumin interneurons, principal neurons, and brain oscillations: A review. Front. Synaptic Neurosci. 2021, 13, 673210. [Google Scholar] [CrossRef]
  242. Hayani, H.; Song, I.; Dityatev, A. Increased excitability and reduced excitatory synaptic input into fast-spiking Ca2 interneurons after enzymatic attenuation of extracellular matrix. Front. Cell. Neurosci. 2018, 12, 149. [Google Scholar] [CrossRef] [PubMed]
  243. Carulli, D.; Rhodes, K.E.; Brown, D.J.; Bonnert, T.P.; Pollack, S.J.; Oliver, K.; Strata, P.; Fawcett, J.W. Composition of perineuronal nets in the adult rat cerebellum and the cellular origin of their components. J. Comp. Neurol. 2006, 494, 559–577. [Google Scholar] [CrossRef]
  244. Lafarga, M.; Berciano, M.T.; Blanco, M. The perineuronal net in the fastigial nucleus of the rat cerebellum. A Golgi and quantitative study. Anat. Embryol. 1984, 170, 79–85. [Google Scholar] [CrossRef] [PubMed]
  245. Schwartz, I.R. Axonal endings in the cat medial superior olive: Coated vesicles and intercellular substance. Brain Res. 1972, 46, 187–202. [Google Scholar] [CrossRef] [PubMed]
  246. Tsien, R.Y. Very long-term memories may be stored in the pattern of holes in the perineuronal net. Proc. Natl. Acad. Sci. USA 2013, 110, 12456–12461. [Google Scholar] [CrossRef] [PubMed]
  247. Rudolph, S.; Badura, A.; Lutzu, S.; Pathak, S.S.; Thieme, A.; Verpeut, J.L.; Wagner, M.J.; Yang, Y.M.; Fioravante, D. Cognitive-affective functions of the cerebellum. J. Neurosci. 2023, 43, 7554–7564. [Google Scholar] [CrossRef] [PubMed]
  248. Frischknecht, R.; Heine, M.; Perrais, D.; Seidenbecher, C.I.; Choquet, D.; Gundelfinger, E.D. Brain extracellular matrix affects AMPA receptor lateral mobility and short-term synaptic plasticity. Nat. Neurosci. 2009, 12, 897–904. [Google Scholar] [CrossRef]
  249. Favuzzi, E.; Marques-Smith, A.; Deogracias, R.; Winterflood, C.M.; Sánchez-Aguilera, A.; Mantoan, L.; Maeso, P.; Fernandes, C.; Ewers, H.; Rico, B. Activity-dependent gating of parvalbumin interneuron function by the perineuronal net protein brevican. Neuron 2017, 95, 639–655.e610. [Google Scholar] [CrossRef]
  250. Jakovljevic, A.; Tucic, M.; Blazikova, M.; Korenic, A.; Missirlis, Y.; Stamenkovic, V.; Andjus, P. Structural and functional modulation of perineuronal nets: In search of important players with highlight on tenascins. Cells 2021, 10, 1345. [Google Scholar] [CrossRef]
  251. Burket, J.A.; Webb, J.D.; Deutsch, S.I. Perineuronal nets and metal cation concentrations in the microenvironments of fast-spiking, parvalbumin-expressing gabaergic interneurons: Relevance to neurodevelopment and neurodevelopmental disorders. Biomolecules 2021, 11, 1235. [Google Scholar] [CrossRef] [PubMed]
  252. Vigetti, D.; Andrini, O.; Clerici, M.; Negrini, D.; Passi, A.; Moriondo, A. Chondroitin sulfates act as extracellular gating modifiers on voltage-dependent ion channels. Cell. Physiol. Biochem. 2008, 22, 137–146. [Google Scholar] [CrossRef] [PubMed]
  253. Ali, A.B.; Islam, A.; Constanti, A. The fate of interneurons, GABA(A) receptor sub-types and perineuronal nets in Alzheimer’s disease. Brain Pathol. 2022, 33, e13129. [Google Scholar] [CrossRef]
  254. Härtig, W.; Derouiche, A.; Welt, K.; Brauer, K.; Grosche, J.; Mäder, M.; Reichenbach, A.; Brückner, G. Cortical neurons immunoreactive for the potassium channel Kv3.1b subunit are predominantly surrounded by perineuronal nets presumed as a buffering system for cations. Brain Res. 1999, 842, 15–29. [Google Scholar] [CrossRef]
  255. Härtig, W.; Singer, A.; Grosche, J.; Brauer, K.; Ottersen, O.P.; Brückner, G. Perineuronal nets in the rat medial nucleus of the trapezoid body surround neurons immunoreactive for various amino acids, calcium-binding proteins and the potassium channel subunit Kv3.1b. Brain Res. 2001, 899, 123–133. [Google Scholar] [CrossRef] [PubMed]
  256. Morawski, M.; Reinert, T.; Meyer-Klaucke, W.; Wagner, F.E.; Troger, W.; Reinert, A.; Jager, C.; Bruckner, G.; Arendt, T. Ion exchanger in the brain: Quantitative analysis of perineuronally fixed anionic binding sites suggests diffusion barriers with ion sorting properties. Sci. Rep. 2015, 5, 16471. [Google Scholar] [CrossRef] [PubMed]
  257. Fawcett, J.W.; Fyhn, M.; Jendelova, P.; Kwok, J.C.F.; Ruzicka, J.; Sorg, B.A. The extracellular matrix and perineuronal nets in memory. Mol. Psychiatry 2022, 27, 3192–3203. [Google Scholar] [CrossRef]
  258. Ross, M.T.; Flores, D.; Bertram, R.; Johnson, F.; Wu, W.; Hyson, R.L. Experience-dependent intrinsic plasticity during auditory learning. J. Neurosci. 2019, 39, 1206–1221. [Google Scholar] [CrossRef]
  259. Song, C.; Ehlers, V.L.; Moyer, J.R., Jr. Trace fear conditioning differentially modulates intrinsic excitability of medial prefrontal cortex-basolateral complex of amygdala projection neurons in infralimbic and prelimbic cortices. J. Neurosci. 2015, 35, 13511–13524. [Google Scholar] [CrossRef]
  260. Sehgal, M.; Song, C.; Ehlers, V.L.; Moyer, J.R. Learning to learn—Intrinsic plasticity as a metaplasticity mechanism for memory formation. Neurobiol. Learn. Mem. 2013, 105, 186–189. [Google Scholar] [CrossRef]
  261. Bekisz, M.; Garkun, Y.; Wabno, J.; Hess, G.; Wrobel, A.; Kossut, M. Increased excitability of cortical neurons induced by associative learning: An ex vivo study. Eur. J. Neurosci. 2010, 32, 1715–1725. [Google Scholar] [CrossRef] [PubMed]
  262. Lorenzetti, F.D.; Mozzachiodi, R.; Baxter, D.A.; Byrne, J.H. Classical and operant conditioning differentially modify the intrinsic properties of an identified neuron. Nat. Neurosci. 2006, 9, 17–19. [Google Scholar] [CrossRef]
  263. Alkon, D.L. Calcium-mediated reduction of ionic currents: A biophysical memory trace. Science 1984, 226, 1037–1045. [Google Scholar] [CrossRef] [PubMed]
  264. Ruzicka, J.; Dalecka, M.; Safrankova, K.; Peretti, D.; Jendelova, P.; Kwok, J.C.F.; Fawcett, J.W. Perineuronal nets affect memory and learning after synapse withdrawal. Transl. Psychiatry 2022, 12, 480. [Google Scholar] [CrossRef]
  265. Lasek, A.W.; Chen, H.; Chen, W.Y. Releasing addiction memories trapped in perineuronal nets. Trends Genet. 2018, 34, 197–208. [Google Scholar] [CrossRef] [PubMed]
  266. Paylor, J.W.; Wendlandt, E.; Freeman, T.S.; Greba, Q.; Marks, W.N.; Howland, J.G.; Winship, I.R. Impaired cognitive function after perineuronal net degradation in the medial prefrontal cortex. eNeuro 2018, 5, ENEURO.0253-0218.2018. [Google Scholar] [CrossRef]
  267. Xue, Y.X.; Xue, L.F.; Liu, J.F.; He, J.; Deng, J.H.; Sun, S.C.; Han, H.B.; Luo, Y.X.; Xu, L.Z.; Wu, P.; et al. Depletion of perineuronal nets in the amygdala to enhance the erasure of drug memories. J. Neurosci. 2014, 34, 6647–6658. [Google Scholar] [CrossRef]
  268. Knapska, E.; Lioudyno, V.; Kiryk, A.; Mikosz, M.; Górkiewicz, T.; Michaluk, P.; Gawlak, M.; Chaturvedi, M.; Mochol, G.; Balcerzyk, M.; et al. Reward learning requires activity of matrix metalloproteinase-9 in the central amygdala. J. Neurosci. 2013, 33, 14591–14600. [Google Scholar] [CrossRef]
  269. Wright, J.W.; Brown, T.E.; Harding, J.W. Inhibition of hippocampal matrix metalloproteinase-3 and -9 disrupts spatial memory. Neural Plast. 2007, 2007, 073813. [Google Scholar] [CrossRef] [PubMed]
  270. Sanchez, B.; Kraszewski, P.; Lee, S.; Cope, E.C. From molecules to behavior: Implications for perineuronal net remodeling in learning and memory. J. Neurochem. 2023, 1–23. [Google Scholar] [CrossRef] [PubMed]
  271. Evans, A.; Terstege, D.J.; Scott, G.A.; Tsutsui, M.; Epp, J.R. Neurogenesis mediated plasticity is associated with reduced neuronal activity in CA1 during context fear memory retrieval. Sci. Rep. 2022, 12, 7016. [Google Scholar] [CrossRef] [PubMed]
  272. Lesnikova, A.; Casarotto, P.; Moliner, R.; Fred, S.M.; Biojone, C.; Castren, E. Perineuronal net receptor PTPsigma regulates retention of memories. Front. Synaptic Neurosci. 2021, 13, 672475. [Google Scholar] [CrossRef] [PubMed]
  273. Thompson, E.H.; Lensjø, K.K.; Wigestrand, M.B.; Malthe-Sørenssen, A.; Hafting, T.; Fyhn, M. Removal of perineuronal nets disrupts recall of a remote fear memory. Proc. Natl. Acad. Sci. USA 2018, 115, 607–612. [Google Scholar] [CrossRef]
  274. Banerjee, S.B.; Gutzeit, V.A.; Baman, J.; Aoued, H.S.; Doshi, N.K.; Liu, R.C.; Ressler, K.J. Perineuronal nets in the adult sensory cortex are necessary for fear learning. Neuron 2017, 95, 169–179. [Google Scholar] [CrossRef]
  275. Morikawa, S.; Ikegaya, Y.; Narita, M.; Tamura, H. Activation of perineuronal net-expressing excitatory neurons during associative memory encoding and retrieval. Sci. Rep. 2017, 7, 46024. [Google Scholar] [CrossRef]
  276. Hylin, M.J.; Orsi, S.A.; Moore, A.N.; Dash, P.K. Disruption of the perineuronal net in the hippocampus or medial prefrontal cortex impairs fear conditioning. Learn. Mem. 2013, 20, 267–273. [Google Scholar] [CrossRef]
  277. Karpova, N.N.; Pickenhagen, A.; Lindholm, J.; Tiraboschi, E.; Kulesskaya, N.; Ágústsdóttir, A.; Antila, H.; Popova, D.; Akamine, Y.; Sullivan, R.; et al. Fear erasure in mice requires synergy between antidepressant drugs and extinction training. Science 2011, 334, 1731–1734. [Google Scholar] [CrossRef] [PubMed]
  278. Gogolla, N.; Caroni, P.; Luthi, A.; Herry, C. Perineuronal nets protect fear memories from erasure. Science 2009, 325, 1258–1261. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Rat eyeblink conditioning. A waterfall plot of eyeblink responding in a rat shown as filtered and rectified EMG signals (in arbitrary units) across 100 (Trial Number) stimulus presentations of a 380-ms tone conditioned stimulus (CS, gray rectangle starting 200 ms from trial onset) paired 90 times with a 100-ms periorbital shock unconditioned stimulus (US). The tone CS was presented alone every tenth trial. The EMG signal was blanked at the break in the x axis to prevent the low-voltage EMG signal from being swamped by the larger voltage of the shock, which would otherwise have produced ringing in the EMG amplifier that would have outlasted the response. Part of the unconditioned response can be seen after the break in the EMG signal. Unpublished data.
Figure 1. Rat eyeblink conditioning. A waterfall plot of eyeblink responding in a rat shown as filtered and rectified EMG signals (in arbitrary units) across 100 (Trial Number) stimulus presentations of a 380-ms tone conditioned stimulus (CS, gray rectangle starting 200 ms from trial onset) paired 90 times with a 100-ms periorbital shock unconditioned stimulus (US). The tone CS was presented alone every tenth trial. The EMG signal was blanked at the break in the x axis to prevent the low-voltage EMG signal from being swamped by the larger voltage of the shock, which would otherwise have produced ringing in the EMG amplifier that would have outlasted the response. Part of the unconditioned response can be seen after the break in the EMG signal. Unpublished data.
Biology 13 00200 g001
Figure 2. Simplified cerebellar circuit involved in eyeblink conditioning. Auditory information from the tone conditioned stimulus (CS) shown in blue travels along mossy fibers (MF) from the pontine nuclei (PN). MFs send collateral fibers to neurons in the anterior interpositus nucleus (AIN) of the deep cerebellar nuclei, and then travel up to the cerebellar cortex and synapse onto granule cells (GC). Granule cell axons then bifurcate and synapse on the dendrites of Purkinje cells (PC). There may be thousands of PF synapses onto one PC. Sensory information from the air puff unconditioned stimulus (US) is shown in red travels from the inferior olive (IO) as climbing fibers (CF) that also send collaterals to the AIN, and then travel up to the cerebellar cortex to wrap around and synapse onto the dendritic tree of a Purkinje cell. There is typically only one climbing fiber for each PC. MF and CF synapses are excitatory and PC output to the AIN is inhibitory (−), as are interneurons in the AIN. Figure from [82].
Figure 2. Simplified cerebellar circuit involved in eyeblink conditioning. Auditory information from the tone conditioned stimulus (CS) shown in blue travels along mossy fibers (MF) from the pontine nuclei (PN). MFs send collateral fibers to neurons in the anterior interpositus nucleus (AIN) of the deep cerebellar nuclei, and then travel up to the cerebellar cortex and synapse onto granule cells (GC). Granule cell axons then bifurcate and synapse on the dendrites of Purkinje cells (PC). There may be thousands of PF synapses onto one PC. Sensory information from the air puff unconditioned stimulus (US) is shown in red travels from the inferior olive (IO) as climbing fibers (CF) that also send collaterals to the AIN, and then travel up to the cerebellar cortex to wrap around and synapse onto the dendritic tree of a Purkinje cell. There is typically only one climbing fiber for each PC. MF and CF synapses are excitatory and PC output to the AIN is inhibitory (−), as are interneurons in the AIN. Figure from [82].
Biology 13 00200 g002
Figure 3. Development of the perineuronal net in the anterior interpositus nucleus of the rat cerebellum. The top panels show WFA reactivity (red), DAPI (4′,6-diamidino-2-phenylindole, blue), and MAP2 (microtubule-associated protein 2, green) reactivity in the rat AIN at P12 (A), P18 (B), and P30 (C) at 20×. Scale bars = 100 μm. The bottom panels show an increase in WFA reactivity (red) alone at P12, P30, and P90 at 63×. Scale bars = 50 μm. Figure modified from O’Dell et al. [201].
Figure 3. Development of the perineuronal net in the anterior interpositus nucleus of the rat cerebellum. The top panels show WFA reactivity (red), DAPI (4′,6-diamidino-2-phenylindole, blue), and MAP2 (microtubule-associated protein 2, green) reactivity in the rat AIN at P12 (A), P18 (B), and P30 (C) at 20×. Scale bars = 100 μm. The bottom panels show an increase in WFA reactivity (red) alone at P12, P30, and P90 at 63×. Scale bars = 50 μm. Figure modified from O’Dell et al. [201].
Biology 13 00200 g003
Figure 4. Disruption of the PNN in the DCN with ChABC. The left panel shows WFA labeling in the left anterior interpositus of the deep cerebellar nuclei four days after the rat received an infusion of the enzyme chondroitinase ABC (ChABC). The right panel shows the right anterior interpositus nucleus in the same rat that received an infusion of saline. ChABC was found to have reduced the number of WFA-labeled neurons (42.38 ± 5.24%) compared to the side receiving the vehicle (68.78 ± 5.14%), p < 0.0001. Arrowheads indicate WFA-labeled neurons.
Figure 4. Disruption of the PNN in the DCN with ChABC. The left panel shows WFA labeling in the left anterior interpositus of the deep cerebellar nuclei four days after the rat received an infusion of the enzyme chondroitinase ABC (ChABC). The right panel shows the right anterior interpositus nucleus in the same rat that received an infusion of saline. ChABC was found to have reduced the number of WFA-labeled neurons (42.38 ± 5.24%) compared to the side receiving the vehicle (68.78 ± 5.14%), p < 0.0001. Arrowheads indicate WFA-labeled neurons.
Biology 13 00200 g004
Figure 5. Increase in afterhyperpolarization of a neuron in the DCN after incubation in ChABC. An action potential recorded in a principal neuron of the rat AIN following incubation of a slice of the cerebellum in chondroitinase ABC (ChABC). Slices were incubated for 8 h in either a ChABC concentration of 0.25 U/mL (red trace, ChABC) or 250 μL of the 0.01% bovine serum albumin solution added to the ACSF as a vehicle (black trace, Control). The figure shows that after ChABC incubation, the size of the afterhyperpolarization (AHP) was significantly larger. The statistical results and methodological details are reported in [201].
Figure 5. Increase in afterhyperpolarization of a neuron in the DCN after incubation in ChABC. An action potential recorded in a principal neuron of the rat AIN following incubation of a slice of the cerebellum in chondroitinase ABC (ChABC). Slices were incubated for 8 h in either a ChABC concentration of 0.25 U/mL (red trace, ChABC) or 250 μL of the 0.01% bovine serum albumin solution added to the ACSF as a vehicle (black trace, Control). The figure shows that after ChABC incubation, the size of the afterhyperpolarization (AHP) was significantly larger. The statistical results and methodological details are reported in [201].
Biology 13 00200 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Schreurs, B.G.; O’Dell, D.E.; Wang, D. The Role of Cerebellar Intrinsic Neuronal Excitability, Synaptic Plasticity, and Perineuronal Nets in Eyeblink Conditioning. Biology 2024, 13, 200. https://doi.org/10.3390/biology13030200

AMA Style

Schreurs BG, O’Dell DE, Wang D. The Role of Cerebellar Intrinsic Neuronal Excitability, Synaptic Plasticity, and Perineuronal Nets in Eyeblink Conditioning. Biology. 2024; 13(3):200. https://doi.org/10.3390/biology13030200

Chicago/Turabian Style

Schreurs, Bernard G., Deidre E. O’Dell, and Desheng Wang. 2024. "The Role of Cerebellar Intrinsic Neuronal Excitability, Synaptic Plasticity, and Perineuronal Nets in Eyeblink Conditioning" Biology 13, no. 3: 200. https://doi.org/10.3390/biology13030200

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop