Abstract
The quest to understand how the brain is able to store information for later retrieval has been pursued by many scientists through the years. Although many have made very significant contributions to the field and our current understanding of the process, few have played as pivotal a role in advancing our understanding as William T. Greenough. The current report will utilize associative learning, a training paradigm that has greatly assisted in our understanding of memory consolidation, to demonstrate how findings emerging from the Greenough laboratory helped to not only shape our current understanding of learning induced anatomical plasticity, but to also launch future analyses into the molecular players involved in this process, especially the Fragile X Mental Retardation Protein.
Keywords: eyeblink conditioning, synapse, learning, synaptic plasticity, multiple synapse bouton, perforated, synapses
INTRODUCTION
The process by which we are able to form a memory for later retrieval has been a topic of discussion and subsequent experimentation for decades. Some of the oldest and best-characterized behavioral tasks used in laboratories for examining learning substrates are associative conditioning tasks. Typically, a neutral conditioned stimulus (CS) is paired with a salient unconditioned stimulus (US) that evokes an unconditioned response (UR). After repeated CS–US pairings, in which the CS consistently predicts the onset of the US, the CS gains the ability to elicit a learned conditioned response (CR). Studies have demonstrated that the temporal relationship between the CS and US dictates the involvement of different brain regions and neural circuits. For example, in delay eyeblink conditioning the US either immediately follows or co-terminates with the CS. Thus there is no separation in time between the two stimuli. This form of learning is mediated by brainstem-cerebellar processing (Clark, McCormick, Lavond, & Thompson, 1984; Mauk & Thompson, 1987) and is not dependent upon neocortical processing (Norman, Buchwald, & Villablanca, 1977; Oakley & Russell, 1977; Mauk & Thompson, 1987).
TRACE EYEBLINK ASSOCIATIVE LEARNING
In trace eyeblink conditioning, where the CS and US are temporally separated by a stimulus free interval, successful learning requires not only the brainstem-cerebellum circuitry, but also hippocampal and neocortical regions (Fig. 1). The ability to learn the trace association is abolished in humans, rabbits, rats and mice after lesions in the hippocampus and specific regions of the neocortex (Clark & Squire, 1998; Han et al., 2003; Kim, Clark, & Thompson, 1995; Kronforst-Collins & Disterhoft, 1998; McGlinchey-Berroth, Carrillo, Gabrieli, Brawn, & Disterhoft, 1997; McLaughlin, Skaggs, Churchwell, & Powell, 2002; Moyer, Deyo, & Disterhoft, 1990; Solomon, Vander Schaaf, Thompson, & Weisz, 1986; Takehara, Kawahara, & Kirino, 2003; Takehara, Kawahara, Takatsuki, & Kirino, 2002; Tseng, Guan, Disterhoft, & Weiss, 2004; Weible, McEchron, & Disterhoft, 2000; Weiss, Bouwmeester, Power, & Disterhoft, 1999). Further evidence from our laboratory and others has suggested that the hippocampus is involved early in acquisition and plays a critical role during initial consolidation of the trace association. For example, hippocampal lesions 1 day after acquisition of the trace association abolish subsequent expression of the CR, while similar lesions 30 days after acquiring the task fail to have an impact on response (Kim et al., 1995). Analysis of hippocampal CA1 pyramidal neuronal activity in response to the CS and US during trace eyeblink conditioning demonstrated that conditioning-specific neuronal activity in CA1 precedes CR expression (McEchron & Disterhoft, 1997; Weible, O’Reilly, Weiss, & Disterhoft, 2006), suggesting that the functional properties of neurons in the hippocampus were being altered prior to expression of the behavioral output. Further characterization of hippocampal CA1 pyramidal cells revealed a peak in neuronal activity on the day the subjects began exhibiting CRs (McEchron & Disterhoft, 1997). These studies illustrate not only the hippocampal dependency of this task, but also show some of the changes that occur in the hippocampus during acquisition of a trace association.
FIGURE 1.

Simplified schematic of the connections necessary for trace eyeblink conditioning. The dotted box delineates the structures necessary for delay eyeblink conditioning. Note arrows do not necessary indicate monosynaptic connections.
Learning Induced Biophysical Plasticity
Hippocampal neuronal biophysical properties are also changed during conditioning, in a manner that likely contributes to their enhanced firing rates early in the acquisition of conditioned responses. The post-burst afterhyperpolarization (AHP) and spike frequency accommodation are reduced in CA1 and CA3 hippocampal pyramidal neurons compared to neurons from pseudo conditioned and naive subjects (Fig. 2; Moyer, Power, Thompson, & Disterhoft, 2000; Moyer, Thompson, & Disterhoft, 1996; Thompson, Moyer, & Disterhoft, 1996a). The AHP is mediated by calcium dependent outward potassium currents that serve to regulate the discharge of action potentials during a sustained depolarization. Reducing the AHP increases the number of action potentials in response to a long depolarizing current injection, a function of both a decrease in spike frequency accommodation as well as a shorter post-action potential refractory period. Such reductions decrease the interval between action potentials and increase the total firing rate of the neuron. Several calcium and voltage dependent potassium currents are activated during single or multiple action potentials (Storm, 1990). Our laboratory has demonstrated learning-specific reductions in the voltage- and calcium-dependent BK-mediated fast AHP involved in repolarizing the action potential (Matthews, Weible, Shah, & Disterhoft, 2008; Matthews & Disterhoft, 2009), and in the SK-mediated medium AHP and the slow AHP in CA1 pyramidal neurons following both trace eyeblink conditioning and Morris water maze training (Fig. 3; Matthews, Linardakis, & Disterhoft, 2009; McKay, Matthews, Oliveira, & Disterhoft, 2009; Oh, Kuo, Wu, Sametsky, & Disterhoft, 2003). The conditioning-specific change in the slow AHP is dependent upon PKA for its expression (Oh, McKay, Power, & Disterhoft, 2009), and it is likely that these changes in intrinsic excitability are protein synthesis-dependent and involve CREB activation (Barco, 2007; Oh et al., 2009). More complete descriptions of the intrinsic excitability changes that we and others have observed during and after learning have been reviewed elsewhere (Disterhoft & Oh, 2006).
FIGURE 2.

Trace eyeblink conditioning reduces the after-hyperpolarization (AHP) compared to pseudo conditioned subjects. Figure modified from Figure 3 in Matthews and Disterhoft (2009).
FIGURE 3.

Acquisition on a water maze task reduces hippocampal afterhyperpolarization (AHP). Learners exhibited significantly reduced AHP compared to subjects that were not able to acquire the task (Nonlearners) and Controls. Figure modified from Figure 3 in Oh et al. (2003).
These studies have greatly enhanced our understanding of the neural substrates of hippocampal plasticity during trace-conditioning. Interestingly, other behavioral paradigms have demonstrated similar hippocampal biophysical plasticity. For example, acquisition on a water maze task, a commonly utilized hippocampus dependent spatial learning paradigm, also results in a reduction in the hippocampal AHP (Oh et al., 2003; Tombaugh, Rowe, & Rose, 2005). Furthermore, Green and Greenough (1986) demonstrated that complex rearing, a visual and motor learning paradigm where rodents are typically reared in a large toy-filled environment, exhibited an increase in dentate granule cell population spikes and increased excitatory postsynaptic potential (EPSP) slope compared to normal caged controls. These findings parallel our observations that trace-conditioning increases the slope of the Shaffer collateral-CA1 EPSP compared to neurons from pseudo conditioned and naïve subjects (Power, Thompson, Moyer, & Disterhoft, 1997). Taken together, these studies demonstrate that acquisition of hippocampus dependent tasks alters neuronal properties, resulting in enhanced synaptic transmission and action potential probability. It would be expected that altered synaptic connectivity as reflected in physiological measures would be paralleled in the neuroanatomical characteristics of those same synaptic contacts.
Anatomical Correlates for Learning
The principal postsynaptic component of a synapse is the postsynaptic density (PSD), an electron-dense plate (or plates) lining the cytoplasmic face of the spine or dendrite immediately adjacent to the synaptic cleft and presynaptic bouton. Most synapses, when viewed in serial sections, have a PSD that is continuous, with a shape resembling a disc that can range orders of magnitude in size (Bourne & Harris, 2007; Geinisman, 1993; Nicholson & Geinisman, 2009). Perforated synapses, on the other hand, have at least one discontinuity and can occasionally be composed of multiple, physically separate PSDs synapsing with the same presynaptic bouton (Fig. 4; Bourne & Harris, 2007; Geinisman, 1993; Nicholson & Geinisman, 2009). Though both types of synapses can be found on dendrites and spines, and can be excitatory or inhibitory, most of what is known about these two synaptic subtypes has been determined by studying excitatory axospinous synapses.
FIGURE 4.

Synapses in CA1 stratum radiatum of rat hippocampus. (a1–a3) Serial section conventional electron micrographs showing a multiple synapse bouton (msb) forming two non-perforated axospinous synapses (sp1 and sp2), as well as a single synapse bouton (at) forming a perforated axospinous synapse (sp3). (b1–b4) Serial sections from tissue prepared for post-embedding immunogold electron microscopy for AMPA-type receptors, showing the characteristically high immunoreactivity of perforated synapses. (c1–c3) Serial sections from tissue prepared for post-embedding immunogold electron microscopy for AMPA-type receptors, showing a multiple synapse bouton (msb) that forms two non-perforated axospinous synapses (sp1 and sp2) exhibiting characteristically weak immunoreactivity (white arrowheads) for AMPA-type receptors. In all panels, black arrows delineate the presence of a perforation in the postsynaptic density; black arrowhead delineate the edges of the postsynaptic density; sp, spine; at, axon terminal. Scale bar is 0.5 μm.
In identifying perforated synapses, or subsynaptic plate perforations, as a synaptic correlate of cortical development and experience in an enriched environment, Greenough, West, and DeVoogd (1978) put forth the prescient idea that such synapses may contribute to or underlie experience-dependent synaptic plasticity (reviewed in Geinisman, 2000). As they noted in 1978 (and in their initial descriptions; Peters and Kaiserman-Abramof, 1969, 1970), however, the functional significance of the perforations or subsynaptic plate perforations was less clear. Though their function is not totally understood, our recent work suggests that synapses with perforated PSDs play a central role in influencing neuronal output, regardless of their location. Additionally, our recent work indicates that when the strength of these perforated synapses is decreased, the hippocampus is unable to function normally.
Distance-Dependent Synaptic Scaling
Dendrites are essentially leaky cables in a salt solution. As such, voltage signals that propagate from their site of origin (i.e., the synapse) to the final integration zone in the soma/axon attenuate (roughly) in proportion to the distance between the two locations (reviewed in Rall, 1964; Sjöström, Rancz, Roth, & Haüsser, 2008; Spruston, 2008; Williams & Stuart, 2003). This has long been a source of angst among neuroscientists, since it calls into question the function of excitatory synapses on the most distal dendrites, and implies that, in the absence of compensatory mechanisms, their influence on neuronal output would be less than more proximal synapses. Using dual somatic and dendritic whole-cell patch-clamp recordings, Magee and Cook (2000) solved this long-standing issue by showing that the local unitary EPSPs from synapses on distal dendrites of CA1 pyramidal neurons were much larger than the local unitary EPSPs recorded from synapses located on more proximal dendrites of the same neuron. Remarkably, when these distally- and proximally-generated unitary EPSPs were recorded at the soma, their amplitudes were nearly indistinguishable (~0.2 mV). In other words, synapses increase their strength in proportion to their distance from the soma such that their influence on somatic/axonal depolarization is roughly equivalent. Through a series of elegant experiments, Magee and his group showed that the strength of AMPA-receptor (AMPAR) mediated transmission is enhanced with distance from the soma, a mechanism they termed distance-dependent synaptic scaling (Andrásfalvy & Magee, 2001; Magee & Cook, 2000; Smith, Ellis-Davies, & Magee, 2003).
Whether the strength of all synapses was augmented, or whether only some synaptic subtypes increased their strength with distance from the soma was unknown. One piece of evidence in the support of the latter was the presence of unitary local EPSPs in the distal dendrites that were significantly larger than any found in the proximal recording locations. We and others had shown that perforated synapses had a higher number and density of immunogold particles for AMPARs and NMDA receptors (NMDARs; Baude, Nusser, Molnár, McIlhinney, & Somogyi, 1995; Desmond & Weinberg, 1998; Ganeshina, Berry, Petralia, Nicholson, & Geinisman, 2004a,b), and we hypothesized that a straightforward way for CA1 pyramidal neurons to implement distance-dependent synaptic scaling was to increase the strength or number of perforated synapses with distance from the soma. Using unbiased quantitative serial section electron microscopy in combination with post-embedding immunogold electron microscopy, we found that the number and proportion of perforated synapses increased progressively with distance from the soma in CA1 (Nicholson & Geinisman, 2009; Nicholson et al., 2006). Moreover, AMPAR expression was significantly higher in distal stratum radiatum than in proximal stratum radiatum, whereas NMDAR expression showed no distance-dependent differences. Remarkably, this distance-dependent regulation of (putative) synaptic strength was restricted to perforated synapses. Our colleagues, Nelson Spruston and William Kath along with their students Yael Katz, Rachel Trana, and Vilas Menon, then derived a computational model of a CA1 pyramidal neuron based on these data and showed that the distance-dependent increase in number and strength among perforated synapses could reproduce the findings of Magee and Cook (2000). They found that the amplitude of the simulated EPSPs produced by a conductance change at perforated synapses was larger locally in the distal dendrites than those produced in the proximal dendrites, but both were similar at the soma. Interestingly, the proportion of perforated synapses was highest in the most distal dendritic region, stratum lacunosum-moleculare (SLM), but their AMPAR expression was actually the lowest. The computational model indicated that these synapses, unlike those throughout stratum radiatum, may not utilize unitary EPSPs to drive somatic depolarization. Rather, the data and the computational model suggest that SLM synapses are more likely to act cooperatively to trigger a dendritic spike, which then propagates to the soma/axon to influence neuronal output (see also Ahmed & Siegelbaum, 2009; Golding, Mickus, Katz, Kath, & Spruston, 2005; Golding & Spruston, 1998; Jarsky, Roxin, Kath, & Spruston, 2005; Katz et al., 2009).
As the landmark paper by Greenough et al. (1978) suggested, these peculiar perforated synapses with the subsynaptic plate perforation may be a principal component underlying the construction of cortical circuits in an experience-expectant manner (Greenough, 1984; Grossman, Churchill, Bates, Kleim, & Greenough, 2002; Markham, Black, & Greenough, 2007) in that they permit location-independent information processing to occur in cortical pyramidal neurons, essentially allowing a dendritic democracy to operate where each synapse can contribute to neuronal output (Häusser, 2001). It will be important for future studies to determine how critical this compensatory mechanism is for brain function (e.g., Sanderson et al., 2008), what its developmental timescale is and whether it co-emerges with the ontogeny of hippocampus-dependent learning, whether its establishment and maintenance is experience-dependent, and whether the distance-dependent changes in synaptic morphology are a general rule for cortical pyramidal neurons even in regions that lack electrophysiological evidence for such scaling (e.g., Williams & Stuart, 2002). Though perforated synapses captured the attention of the Greenough laboratory long ago, much remains to be discovered about their function in neuronal integration and their role in supporting behavior and cognition.
Biophysical and Anatomical Age-Related Correlates of Learning Impairments
As mammals age, many of them are able to retain their ability to acquire hippocampus-dependent behaviors like maze learning or trace eyeblink conditioning, but a substantial proportion of them show marked impairments (reviewed in Burke & Barnes, 2010; Disterhoft & Oh, 2006; Gallagher & Rapp, 1997; Wilson, Gallagher, Eichenbaum, & Tanila, 2006). Acquisition of trace-associations are impaired in aged rabbits (Deyo, Straube, & Disterhoft, 1989; Graves & Solomon, 1985; Powell, Buchanan, & Hernandez, 1981; Straube, Deyo, Moyer, & Disterhoft, 1990; Thompson et al., 1996a), aged rats (Knuttinen, Gamelli, Weiss, Power, & Disterhoft, 2001), aged mice (Kishimoto, Suzuki, Kawahara, & Kirino, 2001), and aged humans (Finkbiner & Woodruff-Pak, 1991; Knuttinen, Power, Preston, & Disterhoft, 2001).
Our work suggests that a combination of biophysical and structural synaptic abnormalities contribute to age-related impairments, both of which are consistent with reduced excitability and/or excitation in the hippocampus of aged, cognitively impaired animals. For example, the post-burst afterhyperpolarization (AHP) of CA1 pyramidal neurons, which is reduced in response to trace-conditioning (Fig. 2; Moyer et al., 1996, 2000; Thompson, Moyer, & Disterhoft, 1996b), is increased in aged subjects (Fig. 5; Landfield & Pitler, 1984; Matthews et al., 2009; Moyer, Thompson, Black, & Disterhoft, 1992; Moyer et al., 2000; Oh, Power, Thompson, Moriearty, & Disterhoft, 1999). This increase is evidence of decreased cellular excitability that may contribute to age-related deficits in learning trace-eyeblink associations. Consistent with this notion, we have shown that pharmacological treatments that reduce the post-burst AHP in CA1 pyramidal neurons enhance trace conditioning rate in aging but not young animals (nimodipine, an L-type calcium channel blocker (Deyo et al., 1989; Moyer et al., 1992); metrifonate, a cholinesterase inhibitor (Kronforst-Collins et al., 1997; Oh et al., 1999)). Interestingly, aging animals that are trained and learn have reduced post-burst AHPs as compared to learning impaired animals (Moyer et al., 2000; Tombaugh et al., 2005), which suggests that the ability to modify the channels underlying the AHP contributes in an important manner to the functional reorganization of the hippocampal neural circuit that, presumably, allows learning to occur in young animals and some of their aged counterparts. In this regard, it is important to note that there is specificity to the age-related changes in channel function/plasticity. The BK-mediated voltage- and calcium-dependent fast AHP is changed by learning but not by aging; in the same neurons, the slow post-burst AHP, a calcium-dependent outward potassium current, is affected by both (Matthews et al., 2008).
FIGURE 5.

Age related changes in hippocampal afterhyperpolarization (AHP). Top: The slow AHP in CA1 of aged subjects is significantly larger than in young subjects. Bottom: The fast AHP in CA1 of aged subjects is not significantly different from young subjects Modified from Figure 3 in Matthews and Disterhoft (2009).
It is likely that intrinsic biophysical (i.e., non-synaptic) and synaptic deregulation collude to render the aged hippocampus dysfunctional in some animals. Though much is known regarding the abnormal encoding for spatial and cue-related stimuli in aged rats with learning impairments, as well as the age-related biophysical differences, the synaptic substrates underlying the degraded neuronal processing remain unknown. In collaboration with Michela Gallagher’s laboratory, we behaviorally characterized aged rats on the spatial water maze task (~28 months) and then, 1 month later, analyzed synapse number and size using conventional quantitative serial section electron microscopy (Geinisman et al., 2004; Nicholson, Yoshida, Berry, Gallagher, & Geinisman, 2004). Based on previous electrophysiological studies (reviewed in Barnes, 1994), we predicted that aged rats with impairments in Morris water maze learning would show a loss of axospinous synapses in CA1 stratum radiatum, whereas those with preserved learning would have numbers comparable to young adults. Remarkably, we found that the numbers of both perforated and non-perforated synapses in CA1 stratum radiatum remained constant regardless of age or cognitive ability (Geinisman et al., 2004). Though we found that synaptic numerical density was lower in the aged impaired rats, this was offset by a reduction in tissue shrinkage due to fixation and processing, yielding a total number estimate similar to both the aged unimpaired group as well as the young adult rats.
Given the correlation between synapse size and AMPAR and NMDAR expression (Ganeshina et al., 2004b; Katz et al., 2009; Kharazia & Weinberg, 1999; Nusser et al., 1998; Racca, Stephenson, Streit, Roberts, & Somogyi, 2000), one possible synaptic substrate of hippocampal dysfunction in aged learning-impaired rats, in the absence of frank synapse loss, would be a reduction in size of the PSD of axospinous synapses in CA1 stratum radiatum. Since synapse number was maintained across the lifespan in rats, a reduction in the size of their PSDs would suggest a loss of synaptic receptors. Though subtle in its morphological change, a loss of receptors would reduce EPSP amplitude and perhaps reflect a conversion of functional synapses into silent or very weak ones. Using unbiased serial section electron microscopy, we measured the length of each PSD profile in serial sections for perforated and non-perforated axospinous synapses, and estimated their areas as the product of the cumulative PSD length and section thickness (Nicholson et al., 2004). This approach revealed a remarkably specific correlate of cognitive dysfunction in aged rats: a reduction in the size of perforated axospinous synapses. Though the size and distribution of sizes among the non-perforated synapses for young adult, aged learning-unimpaired, and aged learning-impaired rats were not significantly different from each other, PSD sizes for perforated synapses from aged rats with spatial learning impairments showed a reduction in the prevalence of larger PSDs relative to their age-matched cognitively intact counterparts as well as the young adult group. Interestingly, segmented, completely partitioned (SCP) synapses, a perforated synaptic subtype closely linked to long-term potentiation (Geinisman, 1993; Toni et al., 2001) which show an unusually high expression of synaptic AMPARs (Ganeshina et al., 2004b; Nicholson & Geinisman, 2009), showed reductions in size in both aged groups. Learning impairments were only present, however, when the size reduction among SCP synapses was accompanied by a reduction in the size of the other perforated synaptic subtypes. These data suggest that, rather than a loss of synapses, a reduction in their strength is just as disruptive to hippocampal function, particularly when the hypothesized weakening affects the perforated synapses.
Taken together, our work suggests that perforated synapses are important for maintaining lines of communication between the dendritic arborizations, where inputs are integrated initially, and the soma/axon where these integrated voltage signals are summed into the final output of the neuron in the form of an action potential. Additionally, when these synapses are reduced in strength, but not necessarily number, it appears that hippocampal function and hippocampus-dependent behaviors are impaired. As shown by Greenough et al. (1978), this synaptic subtype emerges with development and increases in number when rats are exposed to an enriched environment in a manner similar to that found after induction of long-term potentiation (Geinisman, 2000). An intriguing possibility is that the emergence and maintenance of perforated synapses are linked tightly to brain function and, ultimately, learning and memory across the lifespan of rats.
STRUCTURAL SYNAPTIC CORRELATES OF LEARNING
Greenough and Bailey (1988) surveyed the literature on structural synaptic correlates of learning and memory in vertebrates and invertebrates (Greenough & Bailey, 1988). Though the acquisition time, the brain region, the species, and the learned behaviors themselves differed, Greenough and Bailey (1988) suggested that there are two main structural synaptic substrates of memory. Short-term memories, they proposed, involve a rearrangement of existing connections in the form of the number or location of neurotransmitter vesicles and/or structural modifications in the morphology of the presynaptic active zone or PSD size/curvature. Long-term memories, they surmised, were associated with both morphological changes and synaptogenesis. As shown by both the previous and ensuing work from the Greenough laboratory, learning a complex motor skill is associated with both synaptogenesis and a proliferation of multiple synapse boutons (MSBs; Black, Isaacs, Anderson, Alcantara, & Greenough, 1990; Federmeier, Kleim, & Greenough, 2002; Kleim, Lussnig, Schwarz, Comery, & Greenough, 1996; Kleim, Vij, Ballard, & Greenough, 1997; Kleim et al., 1998), whereas exposure to an enriched environment is associated with morphological changes in the size or shape of the PSD (e.g., Wesa, Chang, Greenough, & West, 1982; Sirevaag & Greenough, 1985) as well as a proliferation of MSBs (Jones, Klintsova, Kilman, Sirevaag, & Greenough, 1997). Our own work within CA1 stratum radiatum corroborates this pattern, but extends it to the acquisition of another complex motor skill: trace eyeblink conditioning (Geinisman et al., 2000; Geinisman, Berry, Disterhoft, Power, & Van der Zee, 2001).
Experience-Dependent Changes in Synapse Size
Previous work from our laboratory had shown that the field EPSP generated in CA1 stratum radiatum, by stimulating the axons of CA3 pyramidal neurons, was larger 1 hr and 1 day after the last trace eyeblink conditioning session in rabbits (Power et al., 1997). Using unbiased, quantitative serial section electron microscopy, we asked whether the acquisition of this hippocampus-dependent task was associated with a learning-related increase in synapse number and size in CA1 stratum radiatum (Geinisman et al., 2000; Figure 6). In contrast to our own expectations, synapse number was not higher in the trace conditioned group relative to a pseudoconditioned yoked control group, or naïve rabbits. Additionally, there were no changes in size among perforated synapses. Rather, a subtle morphological change restricted to the smallest non-perforated synapses emerged as the synaptic change associated with the acquisition of the trace eyeblink conditioned response. Specifically, we found that trace conditioning resulted in a ~15% reduction in the number of the smallest non-perforated synapses (Fig. 7).
FIGURE 6.

Scenarios in which MSB proliferation might support activity-dependent synaptic reorganization. (A) Experience leads to synaptogenesis by supporting the formation of new axospinous synapses onto axonal terminals that already synapse with a dendritic spine. (B) Synaptic pruning leads to the elimination of some synapses concomitantly with the formation of new ones. (C) Motile spines change their presynaptic terminal to one already synapsing with a dendritic spine.
FIGURE 7.

Analysis of trace eyeblink conditioning on post-synaptic density size and the number of multiple synapse boutons in CA1 of the hippocampus. Figures are modified from figures in Geinisman et al. (2000, 2001).
In CA1, many non-perforated synapses are smaller than even the smallest perforated synapse (Ganeshina et al., 2004b; Nicholson & Geinisman, 2009). Moreover, the expression level for AMPARs and NMDARs is much lower among the non-perforated synapses compared to their perforated counterparts, even when their PSD sizes are identical (Ganeshina et al., 2004a,b). Lastly, many (30–50%) non-perforated synapses lack AMPAR immunoreactivity, and this AMPAR immuno-negativity is most common among the smallest synapses (Ganeshina et al., 2004a,b; Nicholson et al., 2006; Nicholson & Geinisman, 2009; Nusser et al., 1998; Petralia et al., 1999; Racca et al., 2000; Takumi, Ramirez-León, Laake, Rinvik, & Ottersen, 1999). Therefore, one possibility is that the conditioning-specific decrease in the proportion of the smallest non-perforated synapses reflects a transformation of many of the smallest and presumably AMPAR immunonegative synapses into larger ones that express AMPARs as a result of experience (i.e., trace eyeblink conditioning). This subtle change would allow more synapses to contribute unitary potentials to the field EPSP, and could therefore underlie the conditioning-specific enhancement of synaptic transmission found after trace eyeblink conditioning (Power et al., 1997). Thus, as discussed in Greenough and Bailey (1988), this putative functional change in the absence of net synaptogenesis suggests that a remodeling of existing connections may accompany, even transiently, the acquisition of a complex behavior.
Proliferation of MSBs
Most axonal boutons that make excitatory axospinous synapses in hippocampal area CA1 “contact” one spine. Such single-synapse boutons contrast with MSBs, which can make synapses with 2–5 different spines (Nicholson & Geinisman, 2009; Sorra & Harris, 1993). MSBs have been found to increase in number under a variety of conditions that induce synaptic plasticity or reorganization, including reactive or recuperative synaptogenesis (Hatton, 1990; Jones, 1999; Kirov, Sorra, & Harris, 1999; Meshul et al., 2000; Steward, Vinsant, & Davis, 1988), visual deprivation (Friedlander, Martin, & Wassenhove-McCarthy, 1991), hormonal fluctuations (Woolley, Wenzel, & Schwartzkroin, 1996; Yankova, Hart, & Woolley, 2001), motor skill learning (Federmeier et al., 2002), exposure to an enriched environment (Jones et al., 1997), induction of long-term potentiation (Toni, Buchs, Nikonenko, Bron, & Muller, 1999), and neurogenesis (Toni et al., 2007). We examined whether the acquisition of trace eyeblink conditioning in rabbits, which did not result in a net synaptogenesis but did produce morphological changes in (presumably) pre-existing connections, was associated with an increase in the number of MSBs (Fig. 7; Geinisman et al., 2001).
We found that trace eyeblink conditioning increased the number of MSBs relative to a pseudoconditioned group as well as naïve controls by ~20%. Interestingly, however, neither the number of synapses per MSB nor the number of perforated synapses per MSB was changed. Nonetheless, the increase in PSD size among non-perforated synapses (Geinisman et al., 2000) combined with the proliferation of MSBs (Geinisman et al., 2001) does suggest that more synapses, some of which may have been strengthened by trace eyeblink conditioning, will be activated simultaneously (Fig. 6).
A caveat for this interpretation is necessary, however, based on our recent work (Nicholson and Geinisman, 2009). MSBs in hippocampal region CA1 are comprised of three main subtypes (1) those involving only non-perforated synapses, (2) those involving only perforated synapses, and (3) those involving a mix of these two synaptic subtypes. When the AMPAR and NMDAR expression of synapses involved in a MSB was compared with that found among PSDs apposed with a single-synapse bouton, we found that the vast majority of synapses involved in a MSB either all show an unusually low level of receptor expression, or include a synapse with an unusually high level of receptor expression and one or more synapses with unusually low levels of receptor expression. Consequently, a presynaptic action potential that invades a MSB is likely to produce significant depolarization in, at most, only one of the spines. Unfortunately, a re-analysis of the MSB data from the Geinisman et al. (2001) study is not possible, but an intriguing idea for future research is that the strength or size of synapses involved in MSBs is increased as a result of conditioning, and that the experience-dependent proliferation of MSBs is associated with a selective retention of relevant connections and a pruning of redundant or irrelevant ones (Greenough, 1984). In other words, trace eyeblink conditioning may increase the number of MSBs as well as increase the strength of their synapses, perhaps reflecting both a rapid experience-dependent morphological change as well as a more enduring synaptic reorganization (Greenough & Bailey, 1988).
NEOCORTICAL LEARNING INDUCED PLASTICITY
These and other studies alike have elegantly described how the hippocampus, involved in the earliest stages of acquisition of the CR and in its initial consolidation or storage in other brain regions, is altered during this process; however, the final storage site for the trace eyeblink CR does not reside in the hippocampus. Hippocampal lesions 30 days after acquisition of a trace association do not hinder subsequent performance (Kim et al., 1995; Takehara et al., 2002). These and other studies, along with various learning theories, have suggested that one of the most likely candidate locations for long-term-memory storage is the neocortex. To facilitate examining age-related learning induced sensory cortical plasticity, recent work from our laboratory utilizing either delay- or trace-conditioning paradigms have taken advantage of the simple, well characterized whisker somatosensory system as a suitable conditioning stimulus (Das, Weiss, & Disterhoft, 2001; Galvez, Weiss, Weible, & Disterhoft, 2006). In the whisker somatosensory system each whisker on the face relays input via a tri-synaptic pathway from the trigeminal nerve to medullary barrelets, then to thalamic barreloids, and finally to layer IV of the somatosensory barrel cortex in a topographically-organized map of the whisker pad (Gould, 1986; Killackey & Leshin, 1975; Woolsey & Van der Loos, 1970; Woolsey, Welker, & Schwartz, 1975). Greenough and Chang (1988) were the first to characterize the developmental growth and retraction of neurons in somatosensory barrel cortex. This highly organized yet simple neuronal pathway development, facilitated examination of neuronal plasticity and the molecular players involved. In further characterization of the somatosensory barrel cortex Greenough and coworkers determined that this dendritic development was dependent upon 6-hydroxydopamine (Loeb, Chang, & Greenough, 1987) and a protein that played a major role in shaping a large portion of William Greenough’s scientific career, the Fragile X Mental Retardation Protein (FMRP; Galvez, Gopal, & Greenough, 2003). Subsequent analyses determined that FMRP, in addition to regulating dendritic growth and development also played a role in spine maturation in somatosensory barrel cortex (Fig. 8; Galvez & Greenough, 2005), occipital cortex (Comery et al., 1997; Irwin, Galvez, & Greenough, 2000; Irwin et al., 2001, 2002; McKinney, Elisseou, & Greenough, 2002) and hippocampus (Grossman, Elisseou, McKinney, & Greenough, 2006).
FIGURE 8.

Mice with the fragile X mental retardation syndrome have more immature appearing spines compared to controls. Photomicrographs of apical dendrites from wildtype (WT) and a mouse model of the fragile X mental retardation syndrome (FraX) mice. Modified from Figure 1 in Galvez and Greenough (2005).
Utilizing this simple well characterized pathway, with whisker stimulation as a CS, we have recently demonstrated in young rabbits and mice that trace-eyeblink conditioning results in a metabolic expansion of the conditioned barrels compared to non-conditioned barrels on the adjacent hemisphere and to pseudo-conditioned barrels in layer IV of the somatosensory cortex (Fig. 9; Galvez et al., 2006; Galvez, Cua, & Disterhoft, 2009). Importantly, there was no such expansion in barrel size for animals that learned non-forebrain dependent delay eyeblink conditioning (Galvez et al., 2009). These results indicate that the expansion is not merely due to sensory activation of the barrel cortex, but is a result of learning an association mediated by input to this region and thus could be a site of storage for the trace association.
FIGURE 9.

Schematic of the metabolic neocortical barrel representation and trace acquisition curves during normal and barrel neocortical lesion conditions. Gray circles delineate conditioned barrels. During normal learning the number of percent conditioned response (%CR) increases with subsequent days of training. With acquisition of the task, the metabolic neocortical barrel representation increases (large gray circles). When the barrels are lesioned (black shape) prior to training subjects cannot learn the association (non-learner; Galvez et al., 2006, 2007, 2009).
Expanding upon these observations, we subsequently lesioned the barrel cortex, either before or following acquisition of the trace association. These analyses determined that pre- and post-conditioning barrel cortical lesions impaired whisker-trace-eyeblink conditioning acquisition and retrieval respectively (Fig. 9; Galvez, Weible, & Disterhoft, 2007). Further analysis of the eyeblink behavior revealed a higher occurrence of non-appropriately timed blinks in the post-conditioning barrel cortical lesion group, suggesting that the subjects knew they needed to blink, however they had lost the ability to appropriately time the blink (Galvez et al., 2007). These results along with the observed cytochrome oxidase barrel cortical expansion suggest that the barrel cortex (1) is a site of storage for an aspect of the trace association and (2) plays a role in CR timing or signaling. These hypotheses are further supported by subsequent analysis of neuronal activity in layer IV of the barrel cortex during conditioning. Such analyses have demonstrated an increase in CS neuronal activity during acquisition (Galvez et al., unpublished work). Further characterization of these neurons and their neuronal activity during and after acquisition of the trace association will assist in our understanding of the neuronal mechanism involved in memory consolidation. This is a process William Greenough has spent his entire career examining.
The scientific contributions that Professor William Greenough has made and continues to make have been pivotal for our understanding of the process by which our brains are able to learn new information. Furthermore, as evident from the small sampling of studies mentioned in this chapter, there are very few in the scientific community who can claim to have had as active a role in shaping and defining the current theories for learning induced anatomical, electrophysiological, and molecular neuronal plasticity such as William Greenough has done through the years.
References
- Ahmed MS, Siegelbaum SA. Recruitment of N-type Ca(2+) channels during LTP enhances low release efficacy of hippocampal CA1 perforant path synapses. Neuron. 2009;63:372–385. doi: 10.1016/j.neuron.2009.07.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andrásfalvy BK, Magee JC. Distance-dependent increase in AMPA receptor number in the dendrites of adult hippocampal CA1 pyramidal neurons. Journal of Neuroscience. 2001;21:9151–9159. doi: 10.1523/JNEUROSCI.21-23-09151.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barco A. The Rubinstein–Taybi syndrome: Modeling mental impairment in the mouse. Genes, Brain, and Behavior. 2007;6(Suppl 1):32–39. doi: 10.1111/j.1601-183X.2007.00320.x. [DOI] [PubMed] [Google Scholar]
- Barnes CA. Normal aging: Regionally specific changes in hippocampal synaptic transmission. Trends in Neurosciences. 1994;17:13–18. doi: 10.1016/0166-2236(94)90029-9. [DOI] [PubMed] [Google Scholar]
- Baude A, Nusser Z, Molnár E, McIlhinney RA, Somogyi P. High-resolution immunogold localization of AMPA type glutamate receptor subunits at synaptic and non-synaptic sites in rat hippocampus. Neuroscience. 1995;69:1031–1055. doi: 10.1016/0306-4522(95)00350-r. [DOI] [PubMed] [Google Scholar]
- Black JE, Isaacs KR, Anderson BJ, Alcantara AA, Greenough WT. Learning causes synaptogenesis, whereas motor activity causes angiogenesis, in cerebellar cortex of adult rats. Proceedings of the National Academy of Sciences of the United States of America. 1990;87:5568–5572. doi: 10.1073/pnas.87.14.5568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bourne J, Harris KM. Do thin spines learn to be mushroom spines that remember? Current Opinion in Neurobiology. 2007;17:381–386. doi: 10.1016/j.conb.2007.04.009. [DOI] [PubMed] [Google Scholar]
- Burke SN, Barnes CA. Senescent synapses and hippocampal circuit dynamics. Trends in Neurosciences. 2010;33:153–161. doi: 10.1016/j.tins.2009.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clark GA, McCormick DA, Lavond DG, Thompson RF. Effects of lesions of cerebellar nuclei on conditioned behavioral and hippocampal neuronal responses. Brain Research. 1984;291:125–136. doi: 10.1016/0006-8993(84)90658-9. [DOI] [PubMed] [Google Scholar]
- Clark RE, Squire LR. Classical conditioning and brain systems: The role of awareness. Science. 1998;280:77–81. doi: 10.1126/science.280.5360.77. [DOI] [PubMed] [Google Scholar]
- Comery TA, Harris JB, Willems PJ, Oostra BA, Irwin SA, Weiler IJ, Greenough WT. Abnormal dendritic spines in fragile X knockout mice: Maturation and pruning deficits. Proceedings of the National Academy of Sciences of the United States of America. 1997;94:5401–5404. doi: 10.1073/pnas.94.10.5401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Das S, Weiss C, Disterhoft JF. Eyeblink conditioning in the rabbit (Oryctolagus cuniculus) with stimulation of the mystacial vibrissae as a conditioned stimulus. Behavioral Neuroscience. 2001;115:731–736. doi: 10.1037//0735-7044.115.3.731. [DOI] [PubMed] [Google Scholar]
- Desmond NL, Weinberg RJ. Enhanced expression of AMPA receptor protein at perforated axospinous synapses. Neuroreport. 1998;9:857–860. doi: 10.1097/00001756-199803300-00017. [DOI] [PubMed] [Google Scholar]
- Deyo RA, Straube KT, Disterhoft JF. Nimodipine facilitates associative learning in aging rabbits. Science. 1989;243:809–811. doi: 10.1126/science.2916127. [DOI] [PubMed] [Google Scholar]
- Disterhoft JF, Oh MM. Learning, aging, and intrinsic neuronal plasticity. Trends in Neurosciences. 2006;29:587–599. doi: 10.1016/j.tins.2006.08.005. [DOI] [PubMed] [Google Scholar]
- Federmeier KD, Kleim JA, Greenough WT. Learning-induced multiple synapse formation in rat cerebellar cortex. Neuroscience Letters. 2002;332:180–184. doi: 10.1016/s0304-3940(02)00759-0. [DOI] [PubMed] [Google Scholar]
- Finkbiner RG, Woodruff-Pak DS. Classical eyeblink conditioning in adulthood: Effects of age and interstimulus interval on acquisition in the trace paradigm. Psychology and Aging. 1991;6:109–117. doi: 10.1037//0882-7974.6.1.109. [DOI] [PubMed] [Google Scholar]
- Friedlander MJ, Martin KAC, Wassenhove-McCarthy D. Effects of monocular visual deprivation on geniculocortical innervation of area 18 in cat. Journal of Neuroscience. 1991;11:3268–3288. doi: 10.1523/JNEUROSCI.11-10-03268.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gallagher M, Rapp PR. The use of animal models to study the effects of aging on cognition. Annual Review of Psychology. 1997;48:339–370. doi: 10.1146/annurev.psych.48.1.339. [DOI] [PubMed] [Google Scholar]
- Galvez R, Greenough WT. Sequence of abnormal dendritic spine development in primary somatosensory cortex of a mouse model of the fragile X mental retardation syndrome. American Journal of Medical Genetics Part A. 2005;135:155–160. doi: 10.1002/ajmg.a.30709. [DOI] [PubMed] [Google Scholar]
- Galvez R, Gopal AR, Greenough WT. Somatosensory cortical barrel dendritic abnormalities in a mouse model of the fragile X mental retardation syndrome. Brain Research. 2003;971:83–89. doi: 10.1016/s0006-8993(03)02363-1. [DOI] [PubMed] [Google Scholar]
- Galvez R, Weible AP, Disterhoft JF. Cortical barrel lesions impair whisker-CS trace eyeblink conditioning. Learning & Memory. 2007;14:94–100. doi: 10.1101/lm.418407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Galvez R, Cua S, Disterhoft JF. Age-related deficits in a forebrain-dependent task, trace-eyeblink conditioning. Neurobiology of Aging. 2009 doi: 10.1016/j.neurobiolaging.2009.11.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Galvez R, Weiss C, Weible AP, Disterhoft JF. Vibrissa-signaled eyeblink conditioning induces somatosensory cortical plasticity. Journal of Neuroscience. 2006;26:6062–6068. doi: 10.1523/JNEUROSCI.5582-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ganeshina O, Berry RW, Petralia RS, Nicholson DA, Geinisman Y. Differences in the expression of AMPA and NMDA receptors between axospinous perforated and nonperforated synapses are related to the configuration and size of postsynaptic densities. Journal of Comparative Neurology. 2004a;468:86–95. doi: 10.1002/cne.10950. [DOI] [PubMed] [Google Scholar]
- Ganeshina O, Berry RW, Petralia RS, Nicholson DA, Geinisman Y. Synapses with a segmented, completely partitioned postsynaptic density express more AMPA receptors than other axospinous synaptic junctions. Neuroscience. 2004b;125:615–623. doi: 10.1016/j.neuroscience.2004.02.025. [DOI] [PubMed] [Google Scholar]
- Geinisman Y. Perforated axospinous synapses with multiple, completely partitioned transmission zones: Probable structural intermediates in synaptic plasticity. Hippocampus. 1993;3:417–433. doi: 10.1002/hipo.450030404. [DOI] [PubMed] [Google Scholar]
- Geinisman Y. Structural synaptic modifications associated with hippocampal LTP and behavioral learning. Cerebral Cortex. 2000;10:952–962. doi: 10.1093/cercor/10.10.952. [DOI] [PubMed] [Google Scholar]
- Geinisman Y, Disterhoft JF, Gundersen HJ, McEchron MD, Persina IS, Power JM, Van der Zee EA, West MJ. Remodeling of hippocampal synapses after hippocampus-dependent associative learning. Journal of Comparative Neurology. 2000;417:49–59. [PubMed] [Google Scholar]
- Geinisman Y, Berry RW, Disterhoft JF, Power JM, Van der Zee EA. Associative learning elicits the formation of multiple-synapse boutons. Journal of Neuroscience. 2001;21:5568–5573. doi: 10.1523/JNEUROSCI.21-15-05568.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Geinisman Y, Ganeshina O, Yoshida R, Berry RW, Disterhoft JF, Gallagher M. Aging, spatial learning, and total synapse number in the rat CA1 stratum radiatum. Neurobiology of Aging. 2004;25:407–416. doi: 10.1016/j.neurobiolaging.2003.12.001. [DOI] [PubMed] [Google Scholar]
- Golding NL, Spruston N. Dendritic sodium spikes are variable triggers of axonal action potentials in hippocampal CA1 pyramidal neurons. Neuron. 1998;21:1189–1200. doi: 10.1016/s0896-6273(00)80635-2. [DOI] [PubMed] [Google Scholar]
- Golding NL, Mickus TJ, Katz Y, Kath WL, Spruston N. Factors mediating powerful voltage attenuation along CA1 pyramidal neuron dendrites. Journal of Physiology. 2005;568:69–82. doi: 10.1113/jphysiol.2005.086793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gould HJ., III Body surface maps in the somatosensory cortex of rabbit. Journal of Comparative Neurology. 1986;243:207–233. doi: 10.1002/cne.902430206. [DOI] [PubMed] [Google Scholar]
- Graves CA, Solomon PR. Age-related disruption of trace but not delay classical conditioning of the rabbit’s nictitating membrane response. Behavioral Neuroscience. 1985;99:88–96. doi: 10.1037//0735-7044.99.1.88. [DOI] [PubMed] [Google Scholar]
- Green EJ, Greenough WT. Altered synaptic transmission in dentate gyrus of rats reared in complex enviroments: Evidence from hippocampal slices maintained in vitro. Journal of Neurophysiology. 1986;55(4):739–750. doi: 10.1152/jn.1986.55.4.739. [DOI] [PubMed] [Google Scholar]
- Greenough WT. Structural correlates of information storage in the mammalian brain: A review and hypothesis. Trends in Neurosciences. 1984;7:229–233. [Google Scholar]
- Greenough WT, Bailey CH. The anatomy of a memory: Convergence of results across a diversity of tests. Trends in Neurosciences. 1988;11:142–147. [Google Scholar]
- Greenough WT, West RW, DeVoogd TJ. Subsynaptic plate perforations: Changes with age and experience in the rat. Science. 1978;202(4372):1096–1098. doi: 10.1126/science.715459. [DOI] [PubMed] [Google Scholar]
- Grossman AW, Churchill JD, Bates KE, Kleim JA, Greenough WT. A brain adaptation view of plasticity: Is synaptic plasticity an overly limited concept? Progress in Brain Research. 2002;138:91–108. doi: 10.1016/S0079-6123(02)38073-7. [DOI] [PubMed] [Google Scholar]
- Grossman AW, Elisseou NM, McKinney BC, Greenough WT. Hippocampal pyramidal cells in adult Fmr1 knockout mice exhibit an immature-appearing profile of dendritic spines. Brain Research. 2006;1084:158–164. doi: 10.1016/j.brainres.2006.02.044. [DOI] [PubMed] [Google Scholar]
- Häusser M. Synaptic function: Dendritic democracy. Current Biology. 2001;11:R10–R12. doi: 10.1016/s0960-9822(00)00034-8. [DOI] [PubMed] [Google Scholar]
- Han CJ, O’Tuathaigh CM, Trigt L, Quinn JJ, Fanselow MS, Mongeau R, Koch C, Anderson DJ. Trace but not delay fear conditioning requires attention and the anterior cingulate cortex. Proceedings of the National Academy of Sciences. 2003;100:13087–13092. doi: 10.1073/pnas.2132313100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hatton GI. Emerging concepts of structure-function dynamics in adult brain: The hypothalamo-neurohypophysial system. Progress in Neurobiology. 1990;34:437–504. doi: 10.1016/0301-0082(90)90017-b. [DOI] [PubMed] [Google Scholar]
- Irwin SA, Galvez R, Greenough WT. Dendritic spine structural anomalies in fragile-X mental retardation syndrome. Cerebral Cortex. 2000;10:1038–1044. doi: 10.1093/cercor/10.10.1038. [DOI] [PubMed] [Google Scholar]
- Irwin SA, Patel B, Idupulapati M, Harris JB, Crisostomo RA, Larsen BP, Kooy F, Willems PJ, Cras P, Kozlowski PB, Swain RA, Weiler IJ, Greenough WT. Abnormal dendritic spine characteristics in the temporal and visual cortices of patients with fragile-X syndrome: A quantitative examination. American Journal of Medical Genetics. 2001;98:161–167. doi: 10.1002/1096-8628(20010115)98:2<161::aid-ajmg1025>3.0.co;2-b. [DOI] [PubMed] [Google Scholar]
- Irwin SA, Idupulapati M, Gilbert ME, Harris JB, Chakravarti AB, Rogers EJ, Crisostomo RA, Larsen BP, Mehta A, Alcantara CJ, Patel B, Swain RA, Weiler IJ, Oostra BA, Greenough WT. Dendritic spine and dendritic field characteristics of layer V pyramidal neurons in the visual cortex of fragile-X knockout mice. American Journal of Medical Genetics. 2002;111:140–146. doi: 10.1002/ajmg.10500. [DOI] [PubMed] [Google Scholar]
- Jarsky T, Roxin A, Kath WL, Spruston N. Conditional dendritic spike propagation following distal synaptic activation of hippocampal CA1 pyramidal neurons. Nature Neuroscience. 2005;8:1667–1676. doi: 10.1038/nn1599. [DOI] [PubMed] [Google Scholar]
- Jones TA. Multiple synapse formation in the motor cortex opposite unilateral sensorimotor cortex lesions in adult rats. Journal of Comparative Neurology. 1999;414:57–66. [PubMed] [Google Scholar]
- Jones TA, Klintsova AY, Kilman VL, Sirevaag AM, Greenough WT. Induction of multiple synapses by experience in the visual cortex of adult rats. Neurobiology of Learning and Memory. 1997;68:13–20. doi: 10.1006/nlme.1997.3774. [DOI] [PubMed] [Google Scholar]
- Katz Y, Menon V, Nicholson DA, Geinisman Y, Kath WL, Spruston N. Synapse distribution suggests a two-stage model of dendritic integration in CA1 pyramidal neurons. Neuron. 2009;63:171–177. doi: 10.1016/j.neuron.2009.06.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kharazia VN, Weinberg RJ. Immunogold localization of AMPA and NMDA receptors in somatic sensory cortex of albino rat. Journal of Comparative Neurology. 1999;412:292–302. doi: 10.1002/(sici)1096-9861(19990920)412:2<292::aid-cne8>3.0.co;2-g. [DOI] [PubMed] [Google Scholar]
- Killackey HP, Leshin S. The organization of specific thalamocortical projections to the posteromedial barrel subfield of the rat somatic sensory cortex. Brain Research. 1975;86:469–472. doi: 10.1016/0006-8993(75)90897-5. [DOI] [PubMed] [Google Scholar]
- Kim JJ, Clark RE, Thompson RF. Hippocampectomy impairs the memory of recently, but not remotely, acquired trace eyeblink conditioned responses. 1995;Behavioral Neuroscience109:195–203. doi: 10.1037//0735-7044.109.2.195. [DOI] [PubMed] [Google Scholar]
- Kirov SA, Sorra KE, Harris KM. Slices have more synapses than perfusion-fixed hippocampus from both young and mature rats. Journal of Neuroscience. 1999;19:2876–2886. doi: 10.1523/JNEUROSCI.19-08-02876.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kishimoto Y, Suzuki M, Kawahara S, Kirino Y. Age-dependent impairment of delay and trace eyeblink conditioning in mice. Neuroreport. 2001;12:3349–3352. doi: 10.1097/00001756-200110290-00040. [DOI] [PubMed] [Google Scholar]
- Kleim JA, Lussnig E, Schwarz ER, Comery TA, Greenough WT. Synaptogenesis and Fos expression in the motor cortex of the adult rat after motor skill learning. Journal of Neuroscience. 1996;16:4529–4535. doi: 10.1523/JNEUROSCI.16-14-04529.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kleim JA, Vij K, Ballard DH, Greenough WT. Learning-dependent synaptic modifications in the cerebellar cortex of the adult rat persist for at least four weeks. Journal of Neuroscience. 1997;17:717–721. doi: 10.1523/JNEUROSCI.17-02-00717.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kleim JA, Swain RA, Armstrong KA, Napper RM, Jones TA, Greenough WT. Selective synaptic plasticity within the cerebellar cortex following complex motor skill learning. Neurobiology of Learning and Memory. 1998;69:274–289. doi: 10.1006/nlme.1998.3827. [DOI] [PubMed] [Google Scholar]
- Knuttinen MG, Power JM, Preston AR, Disterhoft JF. Awareness in classical differential eyeblink conditioning in young and aging humans. Behavioral Neuroscience. 2001;115:747–757. doi: 10.1037//0735-7044.115.4.747. [DOI] [PubMed] [Google Scholar]
- Knuttinen MG, Gamelli AE, Weiss C, Power JM, Disterhoft JF. Age-related effects on eyeblink conditioning in the F344 × BN F1 hybrid rat. Neurobiology of Aging. 2001;22:1–8. doi: 10.1016/s0197-4580(00)00194-9. [DOI] [PubMed] [Google Scholar]
- Kronforst-Collins MA, Disterhoft JF. Lesions of the caudal area of rabbit medial prefrontal cortex impair trace eyeblink conditioning. Neurobiology of Learning and Memory. 1998;69:147–162. doi: 10.1006/nlme.1997.3818. [DOI] [PubMed] [Google Scholar]
- Kronforst-Collins MA, Moriearty PL, Ralph M, Becker RE, Schmidt B, Thompson LT, Disterhoft JF. Metrifonate treatment enhances acquisition of eyeblink conditioning in aging rabbits. Pharmacology, Biochemistry, and Behavior. 1997;56:103–110. doi: 10.1016/S0091-3057(96)00164-5. [DOI] [PubMed] [Google Scholar]
- Landfield PW, Pitler TA. Prolonged Ca2+-dependent afterhyperpolarizations in hippocampal neurons of aged rats. Science. 1984;226:1089–1092. doi: 10.1126/science.6494926. [DOI] [PubMed] [Google Scholar]
- Loeb EP, Chang FL, Greenough WT. Effects of neonatal 6-hydroxydopamine treatment upon morphological organization of the posteromedial barrel subfield in mouse somatosensory cortex. Brain Research. 1987;403:113–120. doi: 10.1016/0006-8993(87)90129-6. [DOI] [PubMed] [Google Scholar]
- Magee JC, Cook EP. Somatic EPSP amplitude is independent of synapse location in hippocampal pyramidal neurons. Nature Neuroscience. 2000;3:895–903. doi: 10.1038/78800. [DOI] [PubMed] [Google Scholar]
- Markham JA, Black JE, Greenough WT. Developmental approaches to the memory process. In: Kesner RP, Martinez JL, editors. Neurobiology of learning and memory. Oxford: Elsevier, Inc.; 2007. pp. 57–102. [Google Scholar]
- Meshul CK, Cogen JP, Cheng HW, Moore C, Krentz L, McNeill TH. Alterations in rat striatal glutamate synapses following a lesion of the cortico- and/or nigrostrial pathway. Experimental Neurology. 2000;165:191–206. doi: 10.1006/exnr.2000.7467. [DOI] [PubMed] [Google Scholar]
- Matthews EA, Disterhoft JF. Blocking the BK channel impedes acquisition of trace eyeblink conditioning. Learning & Memory. 2009;16:106–109. doi: 10.1101/lm.1289809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matthews EA, Linardakis JM, Disterhoft JF. The fast and slow afterhyperpolarizations are differentially modulated in hippocampal neurons by aging and learning. Journal of Neuroscience. 2009;29:4750–4755. doi: 10.1523/JNEUROSCI.0384-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matthews EA, Weible AP, Shah S, Disterhoft JF. The BK-mediated fAHP is modulated by learning a hippocampus-dependent task. Proceedings of the National Academy of Sciences of the United States of America. 2008;105:15154–15159. doi: 10.1073/pnas.0805855105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mauk MD, Thompson RF. Retention of classically conditioned eyelid responses following acute decerebration. Brain Research. 1987;403:89–95. doi: 10.1016/0006-8993(87)90126-0. [DOI] [PubMed] [Google Scholar]
- McEchron MD, Disterhoft JF. Sequence of single neuron changes in CA1 hippocampus of rabbits during acquisition of trace eyeblink conditioned responses. Journal of Neurophysiology. 1997;78:1030–1044. doi: 10.1152/jn.1997.78.2.1030. [DOI] [PubMed] [Google Scholar]
- McGlinchey-Berroth R, Carrillo MC, Gabrieli JD, Brawn CM, Disterhoft JF. Impaired trace eyeblink conditioning in bilateral, medial-temporal lobe amnesia. Behavioral Neuroscience. 1997;111:873–882. doi: 10.1037//0735-7044.111.5.873. [DOI] [PubMed] [Google Scholar]
- McKay BM, Matthews EA, Oliveira FA, Disterhoft JF. Intrinsic neuronal excitability is reversibly altered by a single experience in fear conditioning. Journal of Neurophysiology. 2009;102:2763–2770. doi: 10.1152/jn.00347.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McKinney BC, Elisseou NM, Greenough WT. Dendritic spine abnormalities in Fragile X (FMR-1) knockout mice with a C57BL/6 background. 2002 doi: 10.1002/ajmg.b.30183. SFN Abstract 233.3. [DOI] [PubMed] [Google Scholar]
- McLaughlin J, Skaggs H, Churchwell J, Powell DA. Medial prefrontal cortex and pavlovian conditioning: Trace versus delay conditioning. Behavioral Neuroscience. 2002;116:37–47. [PubMed] [Google Scholar]
- Moyer JR, Deyo RA, Disterhoft JF. Hippocampectomy disrupts trace eye-blink conditioning in rabbits. Behavioral Neuroscience. 1990;104:243–252. doi: 10.1037//0735-7044.104.2.243. [DOI] [PubMed] [Google Scholar]
- Moyer JR, Thompson LT, Disterhoft JF. Trace eyeblink conditioning increases CA1 excitability in a transient and learning-specific manner. Journal of Neuroscience. 1996;16:5536–5546. doi: 10.1523/JNEUROSCI.16-17-05536.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moyer JR, Thompson LT, Black JP, Disterhoft JF. Nimodipine increases excitability of rabbit CA1 pyramidal neurons in an age- and concentration-dependent manner. Journal of Neurophysiology. 1992;68:2100–2109. doi: 10.1152/jn.1992.68.6.2100. [DOI] [PubMed] [Google Scholar]
- Moyer JR, Power JM, Thompson LT, Disterhoft JF. Increased excitability of aged rabbit CA1 neurons after trace eyeblink conditioning. Journal of Neuroscience. 2000;20:5476–5482. doi: 10.1523/JNEUROSCI.20-14-05476.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nicholson DA, Geinisman Y. Axospinous synaptic subtype-specific differences in structure, size, ionotropic receptor expression, and connectivity in apical dendritic regions of rat hippocampal CA1 pyramidal neurons. Journal of Comparative Neurology. 2009;512:399–418. doi: 10.1002/cne.21896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nicholson DA, Trana R, Katz Y, Kath WL, Spruston N, Geinisman Y. Distance-dependent differences in synapse number and AMPA receptor expression in hippocampal CA1 pyramidal neurons. Neuron. 2006;50:431–442. doi: 10.1016/j.neuron.2006.03.022. [DOI] [PubMed] [Google Scholar]
- Nicholson DA, Yoshida R, Berry RW, Gallagher M, Geinisman Y. Reduction in size of perforated postsynaptic densities in hippocampal axospinous synapses and age-related spatial learning impairments. Journal of Neuroscience. 2004;24:7648–7653. doi: 10.1523/JNEUROSCI.1725-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Norman RJ, Buchwald JS, Villablanca JR. Classical conditioning with auditory discrimination of the eye blink in decerebrate cats. Science. 1977;196:551–553. doi: 10.1126/science.850800. [DOI] [PubMed] [Google Scholar]
- Nusser Z, Lujan R, Laube G, Roberts JD, Molnár E, Somogyi P. Cell type and pathway dependence of synaptic AMPA receptor number and variability in the hippocampus. Neuron. 1998;21:545–559. doi: 10.1016/s0896-6273(00)80565-6. [DOI] [PubMed] [Google Scholar]
- Oakley DA, Russell IS. Subcortical storage of Pavlovian conditioning in the rabbit. Physiology & Behavior. 1977;18:931–937. doi: 10.1016/0031-9384(77)90203-7. [DOI] [PubMed] [Google Scholar]
- Oh MM, McKay BM, Power JM, Disterhoft JF. Learning-related postburst afterhyperpolarization reduction in CA1 pyramidal neurons is mediated by protein kinase A. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:1620–1625. doi: 10.1073/pnas.0807708106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oh MM, Power JM, Thompson LT, Moriearty PL, Disterhoft JF. Metrifonate increases neuronal excitability in CA1 pyramidal neurons from both young and aging rabbit hippocampus. Journal of Neuroscience. 1999;19:1814–1823. doi: 10.1523/JNEUROSCI.19-05-01814.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oh MM, Kuo AG, Wu WW, Sametsky EA, Disterhoft JF. Watermaze learning enhances excitability of CA1 pyramidal neurons. Journal of Neurophysiology. 2003;90:2171–2179. doi: 10.1152/jn.01177.2002. [DOI] [PubMed] [Google Scholar]
- Peters A, Kaiserman-Abramof IR. The small pyramidal neuron of the rat cerebral cortex. The synapses upon dendritic spines. Zeitschrift fur Zellforschung und Mikroskopische Anatomie. 1969;100:487–506. doi: 10.1007/BF00344370. [DOI] [PubMed] [Google Scholar]
- Peters A, Kaiserman-Abramof IR. The small pyramidal neuron of the rat cerebral cortex. The perikaryon, dendrites, and spines. American Journal of Anatomy. 1970;127:321–355. doi: 10.1002/aja.1001270402. [DOI] [PubMed] [Google Scholar]
- Petralia RS, Esteban JA, Wang YX, Partridge JG, Ahzo HM, Wenthold RJ, Malinow R. Selective acquisition of AMPA receptors over postnatal development suggests a molecular basis for silent synapses. Nature Neuroscience. 1999;2:31–36. doi: 10.1038/4532. [DOI] [PubMed] [Google Scholar]
- Powell DA, Buchanan SL, Hernandez LL. Age-related changes in classical (Pavlovian) conditioning in the New Zealand albino rabbit. Experimental Aging Research. 1981;7:453–465. doi: 10.1080/03610738108259824. [DOI] [PubMed] [Google Scholar]
- Power JM, Thompson LT, Moyer JR, Disterhoft JF. Enhanced synaptic transmission in CA1 hippocampus after eyeblink conditioning. Journal of Neurophysiology. 1997;78:1184–1187. doi: 10.1152/jn.1997.78.2.1184. [DOI] [PubMed] [Google Scholar]
- Racca C, Stephenson FA, Streit P, Roberts JD, Somogyi P. NMDA receptor content of synapses in stratum radiatum of the hippocampal CA1 area. Journal of Neuroscience. 2000;20:2512–2522. doi: 10.1523/JNEUROSCI.20-07-02512.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rall W. Theoretical significance of dendritic trees for neuronal input–output relations. In: Reiss RF, editor. Neural theory and modeling. Palo Alto: Stanford University Press; 1964. pp. 73–97. [Google Scholar]
- Sanderson DJ, Good MA, Seeburg PH, Sprengel R, Rawlins JN, Bannerman DM. The role of the GluR-A (GluR1) AMPA receptor subunit in learning and memory. Progress in Brain Research. 2008;169:159–178. doi: 10.1016/S0079-6123(07)00009-X. [DOI] [PubMed] [Google Scholar]
- Sirevaag AM, Greenough WT. Differential rearing effects on rat visual cortex synapses. II. Synaptic morphometry. Brain Research. 1985;351:215–226. doi: 10.1016/0165-3806(85)90193-2. [DOI] [PubMed] [Google Scholar]
- Sjöström PJ, Rancz EA, Roth A, Haüsser M. Dendritic excitability and synaptic plasticity. Physiological Reviews. 2008;88:769–840. doi: 10.1152/physrev.00016.2007. [DOI] [PubMed] [Google Scholar]
- Smith MA, Ellis-Davies GC, Magee JC. Mechanism of the distance-dependent scaling of Schaffer collateral synapses in rat CA1 pyramidal neurons. Journal of Physiology. 2003;548:245–258. doi: 10.1113/jphysiol.2002.036376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Solomon PR, Vander Schaaf ER, Thompson RF, Weisz DJ. Hippocampus and trace conditioning of the rabbit’s classically conditioned nictitating membrane response. Behavioral Neuroscience. 1986;100:729–744. doi: 10.1037//0735-7044.100.5.729. [DOI] [PubMed] [Google Scholar]
- Sorra KE, Harris KM. Occurrence and three-dimensional structure of multiple synapses between individual radiatum axons and their target pyramidal cells in hippocampal area CA1. Journal of Neuroscience. 1993;13:3736–3748. doi: 10.1523/JNEUROSCI.13-09-03736.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spruston N. Pyramidal neurons: Dendritic structure and synaptic integration. Nature Reviews. Neuroscience. 2008;9:206–221. doi: 10.1038/nrn2286. [DOI] [PubMed] [Google Scholar]
- Steward O, Vinsant SL, Davis L. The process of reinnervation in the dentate gyrus of adult rats: An ultrastructural study of changes in presynaptic terminals as a result of sprouting. Journal of Comparative Neurology. 1988;267:203–210. doi: 10.1002/cne.902670205. [DOI] [PubMed] [Google Scholar]
- Storm JF. Potassium currents in hippocampal pyramidal cells. Progress in Brain Research. 1990;83:161–187. doi: 10.1016/s0079-6123(08)61248-0. [DOI] [PubMed] [Google Scholar]
- Straube KT, Deyo RA, Moyer JR, Disterhoft JF. Dietary nimodipine improves associative learning in aging rabbits. Neurobiology of Aging. 1990;11:659–661. doi: 10.1016/0197-4580(90)90033-v. [DOI] [PubMed] [Google Scholar]
- Takehara K, Kawahara S, Kirino Y. Time-dependent reorganization of the brain components underlying memory retention in trace eyeblink conditioning. Journal of Neuroscience. 2003;23:9897–9905. doi: 10.1523/JNEUROSCI.23-30-09897.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takehara K, Kawahara S, Takatsuki K, Kirino Y. Time-limited role of the hippocampus in the memory for trace eyeblink conditioning in mice. Brain Research. 2002;951:183–190. doi: 10.1016/s0006-8993(02)03159-1. [DOI] [PubMed] [Google Scholar]
- Takumi Y, Ramirez-León V, Laake P, Rinvik E, Ottersen OP. Different modes of expression of AMPA and NMDA receptors in hippocampal synapses. Nature Neuroscience. 1999;2:618–624. doi: 10.1038/10172. [DOI] [PubMed] [Google Scholar]
- Thompson LT, Moyer JR, Disterhoft JF. Trace eyeblink conditioning in rabbits demonstrates heterogeneity of learning ability both between and within age groups. Neurobiology of Aging. 1996a;17:619–629. doi: 10.1016/0197-4580(96)00026-7. [DOI] [PubMed] [Google Scholar]
- Thompson LT, Moyer JR, Disterhoft JF. Transient changes in excitability of rabbit CA3 neurons with a time course appropriate to support memory consolidation. Journal of Neurophysiology. 1996b;76:1836–1849. doi: 10.1152/jn.1996.76.3.1836. [DOI] [PubMed] [Google Scholar]
- Tombaugh GC, Rowe WB, Rose GM. The slow afterhyperpolarization in hippocampal CA1 neurons covaries with spatial learning ability in aged Fisher 344 rats. Journal of Neuroscience. 2005;25:2609–2616. doi: 10.1523/JNEUROSCI.5023-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Toni N, Buchs PA, Nikonenko I, Bron CR, Muller D. LTP promotes formation of multiple spine synapses between a single axon terminal and a dendrite. Nature. 1999;402:421–425. doi: 10.1038/46574. [DOI] [PubMed] [Google Scholar]
- Toni N, Buchs PA, Nikonenko I, Povilaitite P, Parisi L, Muller D. Remodeling of synaptic membranes after induction of long-term potentiation. Journal of Neuroscience. 2001;21:6245–6251. doi: 10.1523/JNEUROSCI.21-16-06245.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Toni N, Teng EM, Bushong EA, Aimone JB, Zhao C, Consiglio A, van Praag H, Martone ME, Ellisman MH, Gage FH. Synapse formation on neurons born in the adult hippocampus. Nature Neuroscience. 2007;10:727–734. doi: 10.1038/nn1908. [DOI] [PubMed] [Google Scholar]
- Tseng W, Guan R, Disterhoft JF, Weiss C. Trace eyeblink conditioning is hippocampally dependent in mice. Hippocampus. 2004;14:58–65. doi: 10.1002/hipo.10157. [DOI] [PubMed] [Google Scholar]
- Weible AP, McEchron MD, Disterhoft JF. Cortical involvement in acquisition and extinction of trace eyeblink conditioning. Behavioral Neuroscience. 2000;114:1058–1067. doi: 10.1037//0735-7044.114.6.1058. [DOI] [PubMed] [Google Scholar]
- Weible AP, O’Reilly JA, Weiss C, Disterhoft JF. Comparisons of dorsal and ventral hippocampus cornu ammonis region 1 pyramidal neuron activity during trace eye-blink conditioning in the rabbit. Neuroscience. 2006;141:1123–1137. doi: 10.1016/j.neuroscience.2006.04.065. [DOI] [PubMed] [Google Scholar]
- Weiss C, Bouwmeester H, Power JM, Disterhoft JF. Hippocampal lesions prevent trace eyeblink conditioning in the freely moving rat. Behavioural Brain Research. 1999;99:123–132. doi: 10.1016/s0166-4328(98)00096-5. [DOI] [PubMed] [Google Scholar]
- Wesa JM, Chang FL, Greenough WT, West RW. Synaptic contact curvature: Effects of differential rearing on rat occipital cortex. Brain Research. 1982;256:253–257. doi: 10.1016/0165-3806(82)90049-9. [DOI] [PubMed] [Google Scholar]
- Williams SR, Stuart GJ. Dependence of EPSP efficacy on synapse location in neocortical pyramidal neurons. Science. 2002;295(5561):1907–1910. doi: 10.1126/science.1067903. [DOI] [PubMed] [Google Scholar]
- Williams SR, Stuart GJ. Role of dendritic synapse location in the control of action potential output. Trends in Neurosciences. 2003;26:147–154. doi: 10.1016/S0166-2236(03)00035-3. [DOI] [PubMed] [Google Scholar]
- Wilson IA, Gallagher M, Eichenbaum H, Tanila H. Neurocognitive aging: Prior memories hinder new hippocampal encoding. Trends in Neurosciences. 2006;29:662–670. doi: 10.1016/j.tins.2006.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Woolsey TA, Van der Loos H. The structural organization of layer IV in the somatosensory region (SI) of mouse cerebral cortex. The description of a cortical field composed of discrete cytoarchitectonic units. Brain Research. 1970;17:205–242. doi: 10.1016/0006-8993(70)90079-x. [DOI] [PubMed] [Google Scholar]
- Woolsey TA, Welker C, Schwartz RH. Comparative anatomical studies of the SmL face cortex with special reference to the occurrence of “barrels” in layer IV. Journal of Comparative Neurology. 1975;164:79–94. doi: 10.1002/cne.901640107. [DOI] [PubMed] [Google Scholar]
- Woolley CS, Wenzel HJ, Schwartzkroin PA. Estradiol increases frequency of multiple synapse boutons in the hippocampal CA1 region of the adult female rat. Journal of Comparative Neurology. 1996;373:108–117. doi: 10.1002/(SICI)1096-9861(19960909)373:1<108::AID-CNE9>3.0.CO;2-8. [DOI] [PubMed] [Google Scholar]
- Yankova M, Hart SA, Woolley CS. Estrogen increases synaptic connectivity between single presynaptic inputs and multiple postsynaptic CA1 pyramidal cells: A serial electron-microscopic study. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:3525–3530. doi: 10.1073/pnas.051624598. [DOI] [PMC free article] [PubMed] [Google Scholar]
