ArticlePDF Available

Molecular Mechanisms Underlying Memory Consolidation of Taste Information in the Cortex

Frontiers
Frontiers in Behavioral Neuroscience
Authors:

Abstract and Figures

The senses of taste and odor are both chemical senses. However, whereas an organism can detect an odor at a relatively long distance from its source, taste serves as the ultimate proximate gatekeeper of food intake: it helps in avoiding poisons and consuming beneficial substances. The automatic reaction to a given taste has been developed during evolution and is well adapted to conditions that may occur with high probability during the lifetime of an organism. However, in addition to this automatic reaction, animals can learn and remember tastes, together with their positive or negative values, with high precision and in light of minimal experience. This ability of mammalians to learn and remember tastes has been studied extensively in rodents through application of reasonably simple and well defined behavioral paradigms. The learning process follows a temporal continuum similar to those of other memories: acquisition, consolidation, retrieval, relearning, and reconsolidation. Moreover, inhibiting protein synthesis in the gustatory cortex (GC) specifically affects the consolidation phase of taste memory, i.e., the transformation of short- to long-term memory, in keeping with the general biochemical definition of memory consolidation. This review aims to present a general background of taste learning, and to focus on recent findings regarding the molecular mechanisms underlying taste-memory consolidation in the GC. Specifically, the roles of neurotransmitters, neuromodulators, immediate early genes, and translation regulation are addressed.
Molecular mechanisms underlying taste learning in the gustatory cortex: schematic simplified representation assuming the different molecular events take place within the same neuron. Glutamate/neuromodulators (Ach, dopamine) reach the postsynaptic site and affect the neuron through specific receptors (NMDAR, AMPAR, mGluR, muscarinic receptors, dopamine receptors, e.g., D1). The receptors are linked to scaffold proteins that form the postsynaptic density (PSD), e.g., MAGUKs, Shank, Homer. MAGUKs are linked through distal scaffolding proteins (DSP) to CamKII. The complexes of receptors and PSD proteins activate signal transduction and hub molecules, which in turn activate transcription and translation regulation. This chain of events, according to our current understanding, results in a new cellular steady state. It is currently unknown whether all molecular processes depicted occur within a single neuron – a question that remains to be further explored. A, neuromodulators/glutamate; B, receptors; C, scaffold proteins; D, signal transduction and hub molecules; E, translation regulation; F, transcription. Key: neuromodulators – white hexagon; glutamate – white ellipse; receptors of neurotransmitters/neuromodulators – blue; representatives of major protein kinases and phosphatases that may have short-term (CamKIIα) or long-term effects-green; protein translation machinery – purple; transcription factors – coral; immediate early genes – red; second messenger – orange. *CaMKII is known to phosphorylate PSD-95 and SAP-97, but it remains to be clarified whether it phosphorylates PSD-93 and SAP-102 as well. In addition, CamKII is well known to activate cytoplasmic polyadenylation element binding protein (CPEB) in other brain regions and in connection with other forms of learning, thereby affecting protein translation. **The BDNF pathway has been simplified; other proteins participate in this pathway.
… 
This content is subject to copyright.
BEHAVIORAL NEUROSCIENC
E
REVIEW ARTICLE
published: 05 January 2012
doi: 10.3389/fnbeh.2011.00087
Molecular mechanisms underlying memory consolidation
of taste information in the cortex
Shunit Gal-Ben-Ari 1,2 and Kobi Rosenblum1,2*
1Department of Neurobiology, University of Haifa, Haifa, Israel
2Center for Gene Manipulation in the Brain, University of Haifa, Haifa, Israel
Edited by:
Riccardo Brambilla, San Raffaele
Scientific Institute and University, Italy
Reviewed by:
Riccardo Brambilla, San Raffaele
Scientific Institute and University, Italy
Clive R. Bramham, University of
Bergen, Norway
*Correspondence:
Kobi Rosenblum, Department of
Neurobiology and Ethology, Center
for Gene Manipulation in the Brain,
University of Haifa, Mt Carmel, Haifa
31905, Israel.
e-mail: kobir@psy.haifa.ac.il
The senses of taste and odor are both chemical senses. However, whereas an organism
can detect an odor at a relatively long distance from its source, taste serves as the ultimate
proximate gatekeeper of food intake: it helps in avoiding poisons and consuming beneficial
substances. The automatic reaction to a given taste has been developed during evolution
and is well adapted to conditions that may occur with high probability during the lifetime
of an organism. However, in addition to this automatic reaction, animals can learn and
remember tastes, together with their positive or negative values, with high precision and
in light of minimal experience. This ability of mammalians to learn and remember tastes
has been studied extensively in rodents through application of reasonably simple and well
defined behavioral paradigms.The learning process follows a temporal continuum similar to
those of other memories: acquisition, consolidation, retrieval, relearning, and reconsolida-
tion. Moreover, inhibiting protein synthesis in the gustatory cortex (GC) specifically affects
the consolidation phase of taste memory, i.e., the transformation of short- to long-term
memory, in keeping with the general biochemical definition of memory consolidation. This
review aims to present a general background of taste learning, and to focus on recent find-
ings regarding the molecular mechanisms underlying taste–memory consolidation in the
GC. Specifically, the roles of neurotransmitters, neuromodulators, immediate early genes,
and translation regulation are addressed.
Keywords: taste learning, consolidation, gustatory cortex, insular cortex, MAPK, ERK, translation regulation,
conditioned taste aversion
INTRODUCTION
Intake of food and avoidance of poison are crucial to an organ-
ism’s survival in a dynamic environment. The process of deciding
whether food is safe” or “hazardous” relies heavily on the gusta-
tory and somatosensory systems, which are involved in evaluating
diverse properties of putative foods, such as: chemosensory, e.g.,
modality, intensity; orosensory, e.g., texture, temperature, pun-
gency; and gratification capacity (Rosenblum, 2008;Carleton et al.,
2010).
The sense of taste differs from the other senses in two major
characteristics: (1) Each taste input has dual labeling, one related
to its physical (texture and temperature) and chemical features,
and the other to its hedonic value and safety; (2) Deciding on the
safety aspects of a novel sensation, such as the taste of an unfa-
miliar food, involves a different temporal scale from other senses:
taste association with unconditioned stimulus (US) takes 1–12 h
(Bures et al., 1998;Merhav and Rosenblum, 2008), compared with
a few seconds for responses to other sensations.
Two main strategies are used by various species, including
humans, in responding to various tastes: genetic programming,
i.e., affinity to sweet tastes, aversion to bitter ones, and com-
plex learning mechanisms that involve several forebrain structures
(Rosenblum, 2008). Modification of the basic genetic program-
ming and memories of new tastes and their values are expected
to be mediated, at least in part, by the gustatory cortex (GC;
Yamamoto et al., 1984, 1985;Rosenblum, 2008;Doron and Rosen-
blum, 2010), which is defined according to its cytoarchitectonic
boundaries as the dysgranular part of the insular cortex (IC;
Burwell, 2001). This is in line with its unique anatomical connec-
tions (Figure 1), through which it receives multimodal sensory
inputs, including visceral, gustatory, and somatosensory informa-
tion from sensory thalamic nuclei (Fujita et al., 2010). However,
other brain regions, such as the basolateral amygdala (Bla) and
dorsal hippocampus, have been shown to be activated during novel
taste processing (Yefet et al., 2006;Doron and Rosenblum, 2010),
and also to be necessary for safe taste–memory consolidation (De
la Cruz et al., 2008). Most of the studies of the molecular mecha-
nisms underlying taste memory were performed in mice and rats,
therefore, the present review will focus on these studies.
Taste recognition on the receptor level, as well as taste reactiv-
ity, as defined genetically, are beyond the scope of this review; they
have been extensively covered elsewhere (Frank et al., 2008;Car-
leton et al., 2010). The first objective of this review is to describe
taste-related behavior paradigms, in order to elucidate the mole-
cular and cellular mechanisms underlying learning and memory.
We then provide a detailed analysis of the currently known mole-
cular and cellular mechanisms of taste learning in the GC, which
resides in the IC. We aim to focus mainly on studies during the
last decade; previous studies have been reviewed elsewhere (Bures
et al., 1998;Rosenblum, 2008).
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 1
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
FIGURE 1 | Neuroanatomy of the taste system. Chemical stimuli
originating in alimentary sources, upon reaching the oral cavity initiate the
processing of gustatory information (CN, central nucleus; BLA, basolateral
amygdala; NST, nucleus of solitary tract; PBN, parabrachial nucleus; pVPMpc,
parvocellular part of the ventralis postmedial thalamic nucleus of the
thalamus; GC, gustatory cortex). Taste cells, which are broadly tuned to the
diverse taste modalities, are innervated by cranial nerves VII, IX, X, which
project to the primary gustatory nucleus in the brainstem (NST). The NST
sends information to three different systems: the reflex system, the lemniscal
system, and the visceral–limbic system.The reflex system comprises
medullary and reticular-formation neurons, which innervate the cranial motor
nuclei. The lemniscal system consists of projections of the gustatory portion
of the NST to the secondary nucleus situated in the dorsal pons (PBN); this, in
turn, sends axons to the pVPMpc, which ultimately relays gustatory
information to the anterior part of the insular cortex (GC).The visceral–limbic
system refers to a collateral network of connections to the hypothalamus and
limbic areas in the forebrain, which comprises the central gustatory pathway.
The PBN is connected to the amygdala, the hypothalamus, and the
bed-nucleus of the stria terminalis. All limbic gustatory targets are
interconnected with each other as well as with the PBN and the GC.
BEHAVIOR PARADIGMS FOR THE MEASUREMENT OF TASTE
LEARNING, MEMORY, AND CONSOLIDATION
The molecular mechanisms underlying learning and memory are
the subject of ongoing research. Although learning and memory
are considered to be two different processes, they involve a con-
tinuum of events. Following a discrete event, physical changes
underlying memory encoding and processing of the information
to be stored takes place. Ethologically, this is described as the
acquisition phase; it involves creation of an internal representa-
tion of the novel information. This representation remains labile
for some time, while the process of consolidation takes place. Dur-
ing this process the new memory becomes increasingly resistant
to disruption (Alberini, 2011), which can involve several types of
intervention: behavioral (e.g., Merhav et al., 2006), pharmacolog-
ical (e.g., Rosenblum et al., 1993), or others (Bures et al., 1998).
It has been shown that the consolidation of new memory can be
disrupted by many events, including: blocking synthesis of new
RNA or protein, e.g., by actimomycin D or anisomycin, respec-
tively; disruption of the expression or function of specific proteins;
new learning; brain cooling; seizure, e.g., through electric shock;
brain trauma; and regional brain lesions (Alberini, 2011). During
the consolidation phase, the memory is transformed from short-
term memory (STM), which may last from minutes to hours, to
long-term memory (LTM), which may last from days to a lifetime
(McGaugh, 2000;Kandel, 2001;Dudai, 2004). The time frame of
the consolidation phase” can vary within a given learning par-
adigm; it depends on the manipulation, and may reflect several
different cellular and molecular processes (Figure 2). One may
ask whether the difference between consolidation and mainte-
nance of memories is an artificial one, since the processes can also
be regarded as continuous processing of the information by varied
molecular and cellular mechanisms within a certain cortical area.
They can also be considered in terms of temporally differential
involvement of several brain structures.
The next phase of memory processing is use or retrieval of
the memory, during which it is susceptible to further changes,
associated with a process of reconsolidation, which mediates its
restabilization. The stability of a memory depends on its age: new
memories are sensitive to post-reactivation disruption, but older
ones are more resistant (Alberini, 2011).
Ethologically, learning is usually classified into non-associative
(habituation and sensitization), associative (relationships between
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 2
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
FIGURE 2 | The learning process phases: consolidation is
determined by type of disruption. The learning process is
ethologically divided into consecutive phases of acquisition,
consolidation, retrieval, relearning, and reconsolidation. Whereas
acquisition, retrieval, and relearning are defined in positive terms, the
consolidation and reconsolidation phases are defined in negative terms.
Therefore, the duration of the learning process is variable, and
dependent on the type of experimental perturbation used to determine
consolidation time. The different experimental interferences with
consolidation affect different stages of the molecular mechanisms that
underlie learning and memory formation. Although the same types of
interference may be used to disrupt consolidation and reconsolidation, the
time windows of sensitivity to disruption differ between the two phases of
learning. It is therefore unclear whether the molecular processes that occur
during consolidation are identical to the ones that occur during
reconsolidation.
amounts and events), and incidental learning (learning in the
absence of explicit external reinforcement; Gibb et al., 2006;Mor-
ris, 2006;Miranda et al., 2008;Rosenblum, 2008;Lindquist et al.,
2009). Over the years, different behavioral paradigms have been
developed in order to study the different types of learning, and
also to address the abovementioned temporal phases of memory
formation.
One of the most widespread taste-learning paradigms is the
negative-learning, conditioned taste aversion (CTA) paradigm, in
which an association is formed between a novel taste (serving as a
conditioned stimulus CS) and malaise (serving as an uncondi-
tioned stimulus US), resulting in the animal’s subsequent avoid-
ance of the novel food associated with delayed food poisoning
(conditioned response CR; Garcia et al., 1955;Bures et al., 1998).
The acquisition of CTA is subserved by specific brain regions,
including the IC and the amygdala, although their precise role in
CTA is still unclear (Yamamoto et al., 1994;Lamprecht and Dudai,
1996), and it has been demonstrated that induction of neurotoxic
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 3
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
lesions in the IC disrupts acquisition of CTA (Roman et al., 2006;
Roman and Reilly, 2007). Similarly, lesions of the basolateral
region (Bla), but not of the central nucleus of the amygdala, selec-
tively disrupt CTA performance (Nachman and Ashe, 1974;Morris
et al., 1999;St Andre and Reilly, 2007). CTA is commonly used
when a hippocampus-independent form of learning is addressed.
A widely used paradigm of positive-learning is latent inhibition
of CTA (LI-CTA), in which an animal learns that a novel food is
safe, and displays less aversion following CTA than animals sub-
jected to CTA alone (Bures et al., 1998). The LI-CTA paradigm
involves presentation of a novel taste to an animal prior to CTA.
Since this early experience elicits no negative consequences, the
animal displays reduced aversion to the same taste following CTA
(Rosenblum et al., 1993). This modulation of behavior can be
attributed either to reduced strength of the association at the time
of the CTA or to competition during the retrieval phase (Lubow,
1989). LI-CTA is a form of incidental learning that depends on
both the quantity of novel taste consumed and the functionality
of the GC (Rosenblum et al., 1993;Merhav and Rosenblum,2008).
Under certain conditions, CTA can be extinguished (Berman
et al., 2003). Extinction reflects a decrease in the CR in the absence
of reinforcement of a conditioned stimulus. Behavioral evidence
indicates that extinction involves an inhibitory learning mecha-
nism in which the extinguished CR reappears following presenta-
tion of an unconditioned stimulus. However, studies have shown
that the memory was not erased in rats and humans (Lin et al.,
2010).
Taste learning and CTA have several advantages as para-
digms for studying molecular mechanisms underlying learning
and memory. These include: one-trial learning; strong inciden-
tal learning; clear and short learning time; minimal behavioral
manipulation, since the animals can learn in their home cage with
very little interference from other modalities; the sensory input
is clearly defined, and therefore can be quantified in molecular
terms; clearly defined cortical area(s) subserve the learning; and
high reproducibility (Bures et al., 1998;Rosenblum, 2008).
LONG-TERM POTENTIATION IN THE INSULAR CORTEX
The most studied form of neuronal adaptations is Hebbian plastic-
ity, which includes long-term potentiation (LTP), and its recipro-
cal, long-term depression (LTD; Collingridge et al., 2004;Malenka
and Bear, 2004;Feldman, 2009;Pozo and Goda, 2010). Hallmark
features of LTP are that synaptic changes are associative, rapidly
induced, and input specific. These features facilitate reinforce-
ment of active synaptic connections with a given set of sensory
stimuli, thus eliciting an activity-induced increase in synaptic effi-
cacy, which is widely expressed in several pleo- and neocortical
areas. Taken together, these characteristics render LTP an attrac-
tive model for a cellular basis for learning and memory (Bliss and
Collingridge, 1993;Neves et al., 2008;Rosenblum, 2008;Sjostrom
et al., 2008).
Long-term potentiation has been described in several brain
areas, including the hippocampus and the cortex, and in
conjunction with taste learning, (Escobar et al., 1998, 2002;
Ramirez-Lugo et al., 2003;Chen et al., 2006;Alme et al.,
2007;Bramham, 2007;McGeachie et al., 2011;Rodriguez-Duran
et al., 2011). It was demonstrated that tetanic stimulation of
the basolateral amygdaloid nucleus (Bla) induced N-methyl-d-
aspartate (NMDA)-dependent but metabotropic glutamate recep-
tor (mGluR)-independent LTP in the IC (Escobar et al., 1998,
2002;Jones et al., 1999). It is important to note that IC–LTP and
CTA both involve similar molecular mechanisms in the IC, such
as NMDA receptor (NMDAR) dependence, activation of ERK1/2,
and induction of immediate early genes (IEGs), including Zif268,
Fos, Arc, and Homer (Jones et al., 1999;Rodriguez-Duran et al.,
2011).
However, it remains to be more robustly proven that LTP-like
mechanisms in the IC subserve taste learning. It was shown that
induction of LTP in the Bla–IC projection prior to CTA enhanced
CTA retention (Escobar and Bermudez-Rattoni, 2000). In addi-
tion, it was shown that on the one hand, intracortical microinfu-
sion of brain-derived neurotrophic factor (BDNF) induced LTP in
the Bla–IC projection of adult rats (Escobar et al., 2003), and on
the other hand, that intracortical microinfusion of BDNF prior to
CTA training enhanced retention of this task (Castillo et al., 2006).
In a follow-up study, Rodriguez-Duran et al. (2011) have shown
that when the paradigm was reversed, i.e., CTA was performed
prior to LTP induction in the Bla–IC projection, CTA training
prevented the subsequent induction of LTP in the Bla–IC pro-
jection for at least 120 h after CTA training. In addition, they
showed that CTA training produced a persistent change in the
possibility of inducing subsequent LTP in the Bla–IC projection in
a protein-synthesis-dependent manner, and inferred that changes
in the possibility of inducing subsequent synaptic plasticity con-
tributed to the formation and persistence of aversive memories
(Rodriguez-Duran et al., 2011).
Long-term potentiation consists of two phases on the molecu-
lar level: induction, which triggers potentiation, and maintenance,
which sustains the potentiation over time. Many molecules have
been shown to be involved in the induction phase of LTP, but
very few have been implicated in the maintenance phase. Since
the working hypothesis is that LTP is the cellular mechanism
underlying LTM storage, the molecular mechanisms relevant to the
maintenance phase are highly important. To the best of our knowl-
edge, to date, only a single molecule, protein kinase Mζ(PKMζ),
has been found to be necessary for maintenance in both the hip-
pocampus (spatial learning) and the IC (CTA): local injection of
ZIP, a PKMζselective inhibitor, into the hippocampus resulted
in reversal of LTP maintenance in vivo and loss of 1day old spa-
tial memory (Pastalkova et al., 2006). Similarly, its injection into
the GC resulted in reversal of long-term CTA memory in a dose-
dependent manner,whereas other serine/threonine protein kinase
inhibitors are capable of interference with long-term memory for-
mation, but are ineffective once the memory has been established
(Shema et al., 2009, 2011).
Protein kinase Mζis the brain-specific atypical protein kinase C
(PKC) isoform,which unlike full-length PKC isoforms,is a cleaved
form comprising the independent catalytic domain of PKCζ, and
is a second messenger-independent kinase. PKMζis constitutively
active in sustaining LTP maintenance, and studies have shown
that it mediates synaptic potentiation specifically during the late
phase of LTP. LTP induction increases new PKMζsynthesis, leading
to enhanced synaptic transmission. The mechanism underlying
L-LTP and spatial memory maintenance by PKMζis thought
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 4
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
to involve AMPA receptor phosphorylation and trafficking, with
subsequent changes in the amplitude of excitatory postsynaptic
potential (EPSP) concomitant with dendritic translation regulated
by PKMζphosphorylation (and inhibition) of Pin1 (Sacktor, 2008;
Vlachos et al., 2008;Navakkode et al., 2010;Parvez et al., 2010;von
Kraus et al., 2010;Westmark et al., 2010;Mei et al.,2011;Sajikumar
and Korte, 2011).
NEUROTRANSMITTERS IN THE GUSTATORY CORTEX
INVOLVED IN TASTE LEARNING
The working hypothesis is that an organism relies on its sensory
system, specifically, in the present context, the sense of taste, to
create an internal representation of a given physical or chem-
ical stimulus. The sensory information is converted into neu-
ronal activity that subserves the various phases of learning, i.e.,
acquisition, consolidation, and retrieval. Information regarding
physical/chemical properties and the significance of a given taste
reaches the GC via several different neurotransmitters, and elicits
the release of several neurotransmitter systems in the GC, where
relevant receptors are expressed as well. Several molecular changes
in the GC have been found to be correlated with novel taste learn-
ing at various time points after exposure to a novel taste. The
molecules involved include acetylcholine (ACh), dopamine, nora-
drenaline, gamma-aminobutyric acid (GABA), glutamate, and
various neuropeptides (Figure 3;Rosenblum et al., 1995, 1996,
1997;Berman et al., 2000;Belelovsky et al., 2005, 2009;Koh and
Bernstein, 2005;Merhav et al., 2006;Banko et al., 2007;Costa-
Mattioli et al., 2007;Elkobi et al., 2008;Merhav and Rosenblum,
2008;Barki-Harrington et al., 2009b;Doron and Rosenblum,2010;
Sweetat et al., 2011). However, only the muscarinic-cholinergic
and NMDARs have been extensively studied with regard to their
roles in taste–memory acquisition, consolidation, and retention
(Jones et al., 1999;Gutierrez et al., 2003;Nunez-Jaramillo et al.,
2008;Rosenblum, 2008).
GLUTAMATE
Physical and chemical taste information is transferred from the
oral cavity to the cortex via fast neurotransmission, mediated by
the neurotransmitter glutamate, the main excitatory neurotrans-
mitter in the mammalian CNS (Rosenblum, 2008;Rondard et al.,
2011). Since the prominence of a given taste is hypothesized to
be mediated via activation of the neuromodulatory system (e.g.,
Kaphzan et al., 2006),it is possible that the interaction between the
two systems produces a long-term taste–memory trace; it is also
likely that it coincides on specific neurons and probably molecules
that can serve as coincidence detectors of the sensory input and its
meaning (Kaphzan et al., 2006).
The glutamate receptor family comprises four types of
receptors: alpha-amino-3-hydroxy-5-methyl-4-isoxazole propi-
onic acid (AMPA), N-methyl-d-aspartic acid (NMDA), kainate,
and mGluRs. AMPA, NMDA, and kainate receptors are ionotropic
receptors, i.e., they can produce complex fast ion influx-mediated
changes in the neuron, whereas mGluRs produce slow sec-
ond messenger-mediated changes in the neuron by activating
G-protein-coupled receptors (GPCRs; Rondard et al., 2011).
Glutamate and dopamine have been implicated in off-line
processing and memory consolidation following CTA, by means
of in vivo microdialysis and capillary electrophoresis (Guzman-
Ramos et al., 2010). In their study, Guzman-Ramos et al.
(2010) demonstrated the occurrence of an amygdala-dependent
dopamine and glutamate reactivation within the IC, about 45 min
after the stimuli association. Furthermore, blockade of dopamin-
ergic D1 and/or NMDARs before the off-line activity impaired
long- but not STM, which suggests dependence on protein syn-
thesis (Guzman-Ramos et al., 2010). In addition, dopamine and
NMDA can synergistically activate extracellular signal-regulated
kinase (ERK)/Mitogen-activated protein kinase (MAPK) signal-
ing, which is necessary for the formation of long-term taste
memory (Kaphzan et al., 2006, 2007).
Activation of NMDARs has been shown to be necessary for
attenuation of the neophobic taste response in both the IC
(Figueroa-Guzman et al., 2006) and the Bla (Figueroa-Guzman
and Reilly, 2008). Acute microinfusions of MK-801, a non-
competitive NMDAR antagonist, into both brain regions revealed
that although there was no effect on the initial magnitude of the
neophobic response, attenuation of gustatory neophobia was pre-
vented. Similarly, microinfusion of mGlu5-selective antagonist,
3-[2-methyl-1,3-thiazol-4yl)ethynyl]pyridine (MTEP), at 0, 1, or
5μg into the rat IC or Bla prior to CTA resulted in enhanced
CTA performance in the case of the Bla, indicated by robust CTA
followed by slower extinction than in control animals, in a dose-
dependent manner. Interestingly, MTEP microinfusion into the
IC resulted in less robust aversion following CTA, in addition to
enhanced extinction, in a dose-dependent manner (Simonyi et al.,
2009). Previous studies that employed systemic administration
of mGlu antagonists before CTA conditioning have demonstrated
that activation of mGlu5, but not of mGlu1 receptors, was required
for CTA learning (Schachtman et al., 2003). In addition, attenua-
tion of CTA can be attained by microinjectionof a broad-spectrum
mGlu antagonist into the Bla (Yasoshima et al., 2000) or the IC
(Berman et al., 2000).
mGlu5 receptors interact with NMDARs, and the two modu-
late one another’s function in several brain regions in a mutually
positive manner: stimulation of either receptor potentiates the
other (Fowler et al., 2011). The two receptors are physically con-
nected with each other through anchor proteins: mGlu5 receptor
binds Homer proteins (Fagni et al., 2004),NMDAR interacts with
postsynaptic density (PSD)-95, and Homer and PSD-95 can be
clustered by Shank all three of which are PSD proteins (Nais-
bitt et al., 1999;Tu et al., 1999). NMDA and mGlu5 can act
synergistically to activate a number of proteins such as MAPKs,
calcium/calmodulin-dependent protein kinase II (CaMKII), and
CREB (Mao and Wang, 2002;Yang et al., 2004).
A plethora of studies using genetic, pharmacological, physi-
ological, and biochemical approaches have indicated the critical
roles played by mGlu5 and NMDA in both the IC and the hip-
pocampus in acquisition and consolidation, but not retrieval,
of memory of aversive tasks, specifically in avoidance learning
and CTA (Schachtman et al., 2003;Cui et al., 2005;Gravius
et al., 2005;Izquierdo et al., 2006;Simonyi et al., 2007, 2009;
Barki-Harrington et al., 2009a;Nunez-Jaramillo et al., 2010).
In a recent follow-up study employing co-administration of an
mGlu5-receptor-positive allosteric modulator, 3-cyano-N-(1,3-
diphenyl-1H-pyrazol-5-yl) benzamide (CDPPB), and an NMDAR
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 5
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
FIGURE 3 | Molecular mechanisms underlying taste learning
in the gustatory cortex: schematic simplified representation
assuming the different molecular events take place within
the same neuron. Glutamate/neuromodulators (Ach, dopamine)
reach the postsynaptic site and affect the neuron through specific
receptors (NMDAR, AMPAR, mGluR, muscarinic receptors, dopamine
receptors, e.g., D1). The receptors are linked to scaffold proteins that form the
postsynaptic density (PSD), e.g., MAGUKs, Shank, Homer. MAGUKs are
linked through distal scaffolding proteins (DSP) to CamKII.The complexes of
receptors and PSD proteins activate signal transduction and hub molecules,
which in turn activate transcription and translation regulation.This chain of
events, according to our current understanding, results in a new cellular
steady state. It is currently unknown whether all molecular processes
depicted occur within a single neuron a question that remains to be further
explored. A, neuromodulators/glutamate; B, receptors; C, scaffold proteins; D,
signal transduction and hub molecules; E, translation regulation; F,
transcription. Key: neuromodulators white hexagon; glutamate white
ellipse; receptors of neurotransmitters/neuromodulators blue;
representatives of major protein kinases and phosphatases that may have
short-term (CamKIIα) or long-term effects-green; protein translation
machinery purple; transcription factors coral; immediate early genes red;
second messenger orange. *CaMKII is known to phosphorylate PSD-95 and
SAP-97, but it remains to be clarified whether it phosphorylates PSD-93 and
SAP-102 as well. In addition, CamKII is well known to activate cytoplasmic
polyadenylation element binding protein (CPEB) in other brain regions and in
connection with other forms of learning, thereby affecting protein translation.
**The BDNF pathway has been simplified; other proteins participate in this
pathway.
antagonist, MK-801, it was shown that NMDA and mGlu5 inter-
acted functionally in CTA-conditioned rats. Whereas CDPPB
administered by itself prior to the conditioning trial had no effect
on CTA or hippocampus-dependent step-down inhibition, co-
administration of the two compounds resulted in attenuation of
learning deficits in both tasks (Fowler et al., 2011).
There is ample evidence in the literature for the importance of
NMDAR, specifically its regulatory NR2B subunit, in taste learn-
ing. For example, CTA conditioning-induced long-lasting tyro-
sine phosphorylation of NR2B specifically in the IC (Rosenblum
et al., 1995). Furthermore, local administration of (2R)-amino-5-
phosphonovaleric acid (APV), a reversibly selective competitive
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 6
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
inhibitor of NMDAR, to the IC prior to CTA conditioning
impaired CTA memory in a brain-region- and time-dependent
manner (Rosenblum et al., 1995, 1996, 1997). In addition, whereas
exposure to a novel taste, e.g., saccharin, increased phosphoryla-
tion in the NMDAR subunits, repeated doses of saccharin at the
same concentration, which rendered the taste familiar, led to a dra-
matic decrease in the levels of serine phosphorylation of NR2A and
NR2B subunits (Nunez-Jaramillo et al., 2008).
NR2B phosphorylation in the IC is not only correlated with
but is necessary for taste learning, as has been demonstrated
through pharmacological and genetic approaches: local admin-
istration of the tyrosine kinase inhibitor, genistein, to the IC
prevented the increase in phosphorylation of NR2B on tyrosine
1472, and attenuated taste–memory formation (Barki-Harrington
et al., 2009b). Additionally,whereas novel taste exposure has been
recently demonstrated to induce intracellular redistribution of
NR2A and NR2B subunits in the IC (Nunez-Jaramillo et al., 2008),
microinjection of genistein to the IC altered this learning-induced
distribution pattern of NMDAR, highlighting the importance of
NR2B tyrosine phosphorylation after learning in determination
of NMDAR distribution (Barki-Harrington et al., 2009b).
In another study, transgenic (Tg) mice over-expressing the
NR2B subunit specifically in the forebrain (which includes the IC)
were shown to have enhanced CTA, as well as slower extinction
rates, although aversion levels were similar in Tg and wild-type
(Wt) mice 30 days after the CTA conditioning. However, under
the LI-CTA paradigm, the Tg mice did not differ from Wt mice
in their aversion levels, and in a paradigm of two-taste LI-CTA
(second order conditioning, in which the mice are exposed to
both novel and familiar tastes) the Tg mice showed attenuation of
enhanced CTA (Li et al., 2010).
GABA AND ACh
Although the involvement of glutamate receptors in the process-
ing of taste learning in the GC has been extensively studied,
other neurotransmitters have been implicated as well. Novel tastes
have been shown to increase ACh levels in the rat IC, whereas a
familiar taste did not (Miranda et al., 2000). Furthermore, phar-
macological inactivation of the nucleus basalis magnocellularis, a
cholinergic and GABAergic source in the basal forebrain, impaired
CTA acquisition (Miranda and Bermudez-Rattoni, 1999). How-
ever,retrieval was not impaired by this manipulation. In addition,
cholinergic activity mediated by muscarinic receptors in the IC
has been shown to be necessary for acquisition and consolida-
tion of contextual memory of inhibitory avoidance (Miranda and
Bermudez-Rattoni, 2007).
Inhibitory GABAergic interneurons have been recently demon-
strated to be activated in response to novel taste learning in
a layer-specific manner hours after taste learning, which sug-
gests that these neurons are involved not only in acquisition, but
also in off-line processing and consolidation of taste informa-
tion (Doron and Rosenblum, 2010). Electrophysiological studies
in anesthetized rats revealed that excitatory propagation in the IC
was primarily regulated by AMPA and GABAAreceptors (Fujita
et al., 2010). Another electrophysiological study in rat-cortex slices,
which employed multiple-whole-cell patch-clamp recording from
layer V GABAergic interneurons and pyramidal cells of rat IC,
demonstrated that carbachol, a cholinergic agonist, increased the
amplitude of unitary inhibitory postsynaptic currents (uIPSCs)
in interneuron-to-interneuron synapses with higher paired-pulse
ratios. However, the same compound induced dose-dependent
suppression of uIPSCs in fast spiking of pyramidal cell synapses.
This attenuation was mitigated by atropine, a muscarinic ACh
receptor antagonist (Yamamoto et al., 2010), and the authors
inferred that carbachol facilitates GABA release in interneuron
synapses with lower release probability, and cholinergic modula-
tion of GABAergic synaptic transmission is differentially regulated
depending on postsynaptic neuron subtypes.” It is important to
note that neurons projecting from one brain-region involved in
taste learning to another (Figure 1) are inherently excitatory, how-
ever, neurons within a certain brain region may be inhibitory
as well. The roles of inhibitory cortical neurons are yet to be
determined.
POSTSYNAPTIC DENSITY-95
Postsynaptic density-95 (PSD-95), a membrane-associated guany-
late kinase (MAGUK), is the major scaffolding protein in the
excitatory PSD and a potent regulator of synaptic strength (Chen
et al., 2011). It has been recently shown that 3 h following novel
taste learning, expression levels of this protein were specifically
elevated in the GC. This elevation has been shown to be necessary
for acquisition of novel taste memory, but not for its retrieval, and
it has not been observed in response to a familiar taste (Elkobi
et al., 2008). Moreover, there was a correlative increase in PSD-95
association with tyrosine-phosphorylated NR2B following novel
taste learning (Barki-Harrington et al., 2009b). A study concerning
spatial learning, which is both hippocampus- and IC-dependent,
has shown that water-maze training induced a translocation of
NMDARs and PSD-95 to lipid raft membrane microdomains,
concomitant with increased NR2B phosphorylation at tyrosine
1472 in the rat IC (Delint-Ramirez et al., 2008), similarly to
novel taste learning, as mentioned above (Barki-Harrington et al.,
2009b).
PSD-95 and other MAGUK family proteins, SAP-97 and PSD-
93, have been shown to interact with NMDAR, thereby regulating
its function, e.g., all three inhibited NR2B-mediated endocyto-
sis (Lavezzari et al., 2003). The MAGUK family proteins have
been shown to mediate NMDAR clustering and/or trafficking by
association with NMDAR NR2 subunits via their C-terminal glu-
tamate serine (aspartate/glutamate) valine motifs. In addition, the
MAGUK proteins interacted differentially with different NMDAR
subtypes, comprised of differing receptor subunit combination
(Cousins et al., 2008). Specifically, PSD-95 interacted with both
NR2A and NR2B (Kornau et al., 1995),and this interaction could
be modulated by either serine or tyrosine phosphorylation.
For example, CaMKII phosphorylated PSD-95 on serine
residue, causing dissociation of NR2A, but not of NR2B from the
NMDA–PSD-95 complex. In addition, PSD-95 itself functioned
as a negative regulator of the tyrosine kinase Src, for which it
served as a contact point to the NMDAR, thus enabling its reg-
ulation (Kalia et al., 2006). Moreover, phosphorylation of NR2A
and NR2B and their associated proteins by Src or Fyn was found
in some cases to enhance their association with PSD-95 (Rong
et al., 2001;Zalewska et al., 2005). Other tyrosine kinases shown
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 7
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
to interact with PSD-95 include c-Abl and Pyk2: the former regu-
lated synaptic clustering of PSD-95 (de Arce et al., 2010); the latter
underwent PSD-95-induced postsynaptic clustering and activa-
tion (Bartos et al., 2010). It has been suggested that the role of the
PSD-95–NMDA complex is to protect NR2 subunits from under-
going cleavage by calcium-dependent proteases, and thereby to
provide a mechanism for regulating NMDAR expression (Dong
et al., 2004).
THE ROLE OF THE MAPK/ERK PATHWAY IN THE GUSTATORY
CORTEX
Mitogen-activated protein kinases are a family of serine/threonine
kinases implicated in regulation of cell proliferation and differenti-
ation (Seger and Krebs, 1995;Belelovsky et al.,2007). Three major
groups of MAPKs have been identified in mammalian cells: extra-
cellular signal-regulated kinase (ERK), c-jun N-terminal kinase
(JNK), and p38 MAP kinase. High levels of the ERK isoforms,
ERK1 (p44 MAPK) and ERK2 (p42 MAPK), have been detected in
neurons in the mature CNS (Fiore et al., 1993). ERK activity has
been shown to be crucial for several forms of learning and mem-
ory, including fear conditioning, CTA memory, spatial memory,
step-down inhibitory avoidance, and object recognition memory.
In a study in which several pharmacological agents were locally
injected into the rat IC, it was found that NMDARs, mGluRs,
muscarinic, beta-adrenergic, and dopaminergic receptors were
all implicated in acquisition of the new taste memory, but not
in its retrieval, although these neurotransmitter/neuromodulator
systems differed in their role in acquisition. In addition, it was
demonstrated that of all the receptors studied, only NMDA and
muscarinic receptors specifically mediated taste-dependent acti-
vation of ERK1–2, whereas dopaminergic receptors regulated the
basal level of ERK1–2 activation (Berman et al., 2000). Several
studies have demonstrated the differential role of ERK1 and 2 in
cell growth and proliferation and synaptic plasticity. For exam-
ple, ERK1 knockout mice displayed enhancement of striatum-
dependent long-term memory, in conjunction with facilitation of
LTP in the nucleus accumbens, indicating the importance of ERK2
(Mazzucchelli et al., 2002;Vantaggiato et al., 2006).
ERK activation occurs downstream to neurotransmitter release
and activation of the forebrain cholinergic neurons during and
immediately after acquisition of an inhibitory avoidance response.
ERK plays a major role in learning, by promoting cellular integra-
tion of divergent downstream effectors that may trigger differing
responses, depending upon which subsets of scaffolding anchors,
target proteins, and regulatory phosphatases are involved (Giovan-
nini, 2006). MEK–ERK, the upstream kinase of ERK, is a crucial
signal transduction cascade in synaptic plasticity and memory
consolidation (Sweatt, 2001), and its inhibition affected both early
and late phases of LTP in the hippocampus (Rosenblum et al.,
2002).
Extracellular signal-regulated kinase activation was shown to
be correlated with novel taste learning, although the amount of
protein remained unchanged (Berman et al., 1998;Belelovsky
et al., 2005). In addition, the expression of LTP in the GC was
ERK-dependent, and operated in a positive feedback mode (Jones
et al., 1999). In a recent study,it was shown that bilateral injection
of U0126, a specific MEK inhibitor, to the GC prior to learning
resulted in inhibition of ERK1/2, as well as attenuation of CTA
(Rosenblum, 2008) and blockade of learning-induced elevation of
PSD-95 3 h following taste learning (Elkobi et al., 2008).
The various MAPKs are activated within differing time periods
after novel taste learning; ERK activation appears to begin a few
minutes up to 1 h following novel taste consumption (Rosenblum,
2008). It has been shown that novel taste consolidation requires
the elevation of ERK1/2 activity in the IC 20 min after taste con-
sumption. However, JNK1/2 was activated 1 h after novel taste
learning, whereas p38 was not modified at any of the examined
time points (Berman et al., 1998). Interestingly, the time scale of
ERK activation was species-specific; it was shown to differ between
mice and rats (Swank and Sweatt,2001). ERK expression following
taste learning is not only time-restricted, but also space-restricted;
it was activated in the GC, but not in the hippocampus, after
the same length of time following learning (Yefet et al., 2006).
The various MAPKs also differed in BDNF-induced activation:
intrahippocampal microinfusion of BDNF that aimed to induce
LTP resulted in rapid phosphorylation of ERK and p38, but not of
JNK (Ying et al., 2002). These effects were observed in the dentate
gyrus, but not in other examined hippocampal regions, and were
shown to be MEK–ERK-dependent.
Whereas the upstream regulation of ERK in the GC is well
studied, our knowledge regarding its downstream targets remains
fragmentary. MAPK substrate, ELK-1, a transcription factor regu-
lating immediate early expression of genes via the serum response
element (SRE) DNA consensus site (Besnard et al.,2011), has been
shown to be phosphorylated in a time frame similar to that of
ERK activation after novel taste learning, and neurotransmitters
required for induction of LTM are also required for ELK-1 phos-
phorylation in the GC in response to taste learning (Berman et al.,
2003). Further studies are needed to identify other possible targets
of ERK and their mechanisms of action underlying taste processing
in the GC.
THE ROLE OF TRANSLATION REGULATION IN
TASTE–MEMORY CONSOLIDATION
The distinctive biochemical characteristic of memory consolida-
tion is its dependence on synthesis of functional proteins in the
relevant brain regions (Davis and Squire, 1984). Indeed, local
application of anisomycin, a protein-synthesis inhibitor, to specific
brain regions affected CTA in a dose-, site-, and time-dependent
manner (Rosenblum et al., 1993;Meiri and Rosenblum, 1998).
For instance, local application of anisomycin to the GC atten-
uated CTA and taste learning under the LI paradigm (Rosen-
blum et al., 1993); however, the same treatment had no effect
on STM (Houpt and Berlin, 1999), which suggests that short-term
taste memory is independent of protein synthesis. Temporally,
taste learning is sensitive to protein-synthesis inhibitor(s) from
just before learning until up to 100 min afterward (Rosenblum,
2008). In addition, extinction of CTA is dependent on func-
tional protein synthesis in the prefrontal cortex (Akirav et al.,
2006). It is generally accepted that protein translation affects
LTM consolidation by modulation of synaptic strength, since
protein-synthesis inhibitors prevented transformation of early
LTP to late LTP. However, other modifications of intrinsic neuronal
properties also subserve learning-related behavioral changes. For
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 8
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
example, learning-induced reduction in the postburst after hyper-
polarization (AHP) lasts for days after completion of training,
and is implicated in maintenance of learned skills. It has been
recently demonstrated that synaptic activation-induced short-
term postburst AHP reduction in hippocampal and pyramidal
neurons could be transformed to a protein-synthesis-dependent
long-term AHP reduction that persisted for long time periods
(Cohen-Matsliah et al., 2010). Such learning-inducedAHP reduc-
tion has been shown to be maintained by persistent activation
of PKC and ERK in piriform cortex neurons (Seroussi et al.,
2002;Cohen-Matsliah et al., 2007), but has been found to be
CaMKII-independent (Liraz et al., 2009).
Translation consists of three phases: initiation, elongation, and
termination. Of these, the initiation phase is the most tightly
regulated, and is affected by phosphorylation of initiation fac-
tors (IFs) and ribosomal proteins (Proud, 2000). Initiation in
eukaryotes serves as the rate-limiting step in protein synthesis,
and therefore serves as an important target for translational con-
trol (Costa-Mattioli et al., 2009). There is considerable evidence in
the literature that increased initiation results in enhanced learning,
and vice versa. In knockout mice lacking the translation repres-
sor eukaryotic initiation factor 4E-binding protein (4EBP2; Banko
et al., 2007) or with reduced phosphorylation of eukaryotic initi-
ation factor 2α(eIF2α;Costa-Mattioli et al., 2007), and therefore
with enhanced initiation, no differences in taste recognition in
parallel to enhanced CTA learning were observed; furthermore,
other types of learning and plasticity in the consolidation phase
were enhanced as well. Conversely, knockout mice lacking either
S6K1 or S6K2, characterized by reduced initiation rates,exhibited
impaired taste learning (Antion et al., 2008). In a model analo-
gous to CTA in the chick, eukaryotic translation initiation factor
2B (eIF2B) was found to be both correlated with and necessary for
taste–memory consolidation (Tirosh et al., 2007).
The second phase of translation, elongation, requires activ-
ity of elongation factors (EFs). Eukaryotic elongation factor 2
(eEF2) mediated ribosomal translocation (Ryazanov and Davy-
dova, 1989) and was phosphorylated on Thr56 by a spe-
cific Ca2+/calmodulin-dependent kinase. Phosphorylation of this
kinase inhibited its activity and led to general inhibition of protein
synthesis (Nairn and Palfrey,1987). It has been shown that follow-
ing novel taste learning, eEF2 phosphorylation was increased in the
GC, indicating attenuation of translation elongation (Belelovsky
et al., 2005). This finding, which is counter-intuitive, implies that
the situation is more complex than implied by the simple model
of “the more IFs, the better the taste learning” (Rosenblum, 2008).
Nevertheless, there was increased initiation in the same samples,
manifested as increased phosphorylation levels of ERK and S6K1
(Belelovsky et al., 2005).
Therefore, we have proposed a putative mechanism, by which
increased initiation concomitant with decreased elongation in
the same neurons in the GC might increase expression levels of
mRNAs that are poorly initiated. This was reflected, for example,
in the case of phosphorylation of the αsubunit of eIF2 (eIF2 α),
which, in turn, led to suppression of general translation (Hinneb-
usch et al., 2000), concomitant with stimulation of translation of
ATF4 (Lu et al., 2004;Vattem and Wek, 2004),which is required for
late phase LTP and LTM (Bartsch et al., 1995;Chen et al., 2003).
The suggested mechanism could perform a switch-like function in
expressing a specific set of mRNAs within a restricted time window
in a cellular microdomain/s such as the synapse.
MAMMALIAN TARGET OF RAPAMYCIN
Some of the correlative changes that follow novel taste or
CTA paradigms have been observed in proteins that are either
direct or indirect targets of the mammalian target of rapamycin
(mTOR), also known as FKBP-12-rapamycin-associated protein
(FRAP), which consists of two TOR complexes (TORC) that dif-
fer in rapamycin sensitivity. TORC1 mediated rapamycin-sensitive
TOR-shared signaling to the translation machinery, the transcrip-
tion apparatus, and other targets (Loewith et al., 2002;Hay and
Sonenberg, 2004), and its blockade by rapamycin interfered with
translation of specific subpopulations of mRNA (Raught et al.,
2001). Downstream targets of mTOR include ribosomal protein
kinase (S6K1) and EF 1A and 2 (eEF1A and eEF2), which are
mostly involved in ribosome recruitment to mRNA, and regula-
tion of both the initiation and elongation phases of translation
(Hay and Sonenberg, 2004).
There is considerable evidence for the importance of the
mTOR pathway in various forms of synaptic plasticity. In Aplysia,
rapamycin application prevented long-term facilitation (Casadio
et al., 1999), and in the rat hippocampal CA1 region, rapamycin
blocked high-frequency stimulation (HFS) and BDNF-induced
LTP ( Tang et al., 2002). In addition, mTOR-dependent activation
of dendritic S6K1 was shown to be necessary for the induc-
tion phase of protein-synthesis-dependent synaptic plasticity
(Cammalleri et al., 2003).
It has been recently shown that following novel taste learn-
ing, two temporal waves of mTOR activation occurred in the GC
(Belelovsky et al., 2009). Furthermore, it was shown that PSD-95
elevation in the GC 3 h following taste learning, which is neces-
sary for LTM (Elkobi et al., 2008), was prevented following local
application of rapamycin to the GC. Another study has shown, by
means of HFS, an interesting interplay between ERK and mTOR
pathways induced at CA3–CA1 synapses: whereas HFS induced
LTP as well as translational proteins regulated by mTOR, the for-
mer induction was blocked by use of ERK inhibitors (Tsokas et al.,
2007). Moreover, this study showed ERK to be not only correlated
with, but necessary for mTOR stimulation by HFS via interac-
tion with phosphoinositide-dependent kinase 1 (PDK1) and Akt,
which are upstream to mTOR.
Other studies that used genetic and pharmacological
approaches in the CA1 region of the hippocampus showed that
activation of mTORC1 facilitated initiation of protein transla-
tion through phosphorylation and inhibition of eukaryotic initi-
ation factor 4E (eIF4E)-binding proteins (4E-BP), which inhibit
complex formation. Phosphorylation of eIF4E on Ser209 is ERK-
dependent and was closely linked with translation of specific
mRNA subpopulations (Panja et al., 2009). Thus, it is currently
accepted that ERK and mTORC1 synergistically regulate eIF4E and
translation initiation in LTM and synaptic plasticity. It has been
recently shown that Pin1 inhibited protein-synthesis induced by
glutamatergic signaling, and it was suggested that eIF4E and 4E-
BP1/2 mediated an increase in dendritic translation induced by
Pin 1 inhibition (Westmark et al., 2010).
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 9
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
It is important to note that other mechanisms have been impli-
cated in translational control of long-lasting synaptic plasticity and
memory, e.g., micro-RNA regulation (miRNA), and it is highly
likely that they participate, at least in part, in taste learning as well.
However, it remains to be clarified whether this is indeed the case.
IMMEDIATE EARLY GENES AND CONVERGENCE OF CS/US
Immediate early genes, activated transiently and rapidly by vari-
ous cellular stimuli, are regulated at the transcription level. They
provide a first response to these stimuli, in advance of protein
synthesis. Many of the IEGs are transcription factors or other
DNA-binding proteins (Plath et al., 2006). Expression levels of
IEGs have been shown to increase in neuronal populations that
subserve stimulus encoding in response to various kinds of behav-
iors (Campeau et al., 1991;Guzowski et al., 2005;Mattson et al.,
2008;Koya et al.,2009). Several such IEGs have been implicated in
taste memory, including cFOS, Activity-regulated cytoskeleton-
associated protein (Arc/Arg3.1), Homer, and BDNF (Saddoris
et al., 2009;Doron and Rosenblum, 2010), and have been shown
to be elevated following other forms of learning as well. The afore-
mentioned IEGs belong to a subclass termed effector neuronal
IEGs, which mediate NMDAR-regulated phenotypic changes in
the brain (Kaufmann and Worley, 1999). Many of these have been
immunohistochemically detected in dendrites (Lyford et al., 1995;
Tsui et al., 1996).
ARC AND HOMER
Arc and Homer are NMDAR-dependent markers for plasticity
(Guzowski et al., 2001). Arc has been shown to play an impor-
tant role in consolidation of synaptic plasticity and memories
as an effector molecule downstream of many signaling pathways
(Shepherd and Bear, 2011). It mediates synaptic homeostatic scal-
ing of AMPA receptors, via a mechanism of interaction with the
endocytosis machinery which, in turn, regulates AMPA receptor
trafficking (Chowdhury et al., 2006;Plath et al., 2006;Shepherd
et al., 2006). Additionally, Arc affects cytoskeletal dynamics, as local
injection of Arc antisense into the dentate gyrus 2 h following LTP
induction resulted in reversal of LTP, as well as dephosphorylation
of actin depolymerization factor/cofilin, and loss of nascent fila-
mentous actin (F-actin) at synaptic sites (Messaoudi et al., 2007;
Bramham et al., 2010). In turn, these changes are instrumental in
regulation of spine morphology by increasing spine density (Pee-
bles et al., 2010). A recent electrophysiological study has shown
that Arc synthesis was regulated by ERK–MNK signaling dur-
ing LTP consolidation in the dentate gyrus in live rats. Although
mTORC1 is activated following HFS stimulation, its pharmaco-
logical inhibition revealed that it is not essential for Arc synthesis
and LTP (Panja et al., 2009).
Homer1 is a PSD scaffolding protein, involved in the regulation
of synaptic metabotropic receptor function; it has been implicated
in structural changes occurring at synapses during long-lasting
neuronal plasticity and development (Xiao et al., 1998). Both Arc
and Zif268 are required for generation of mRNA-dependent LTP.
Additionally, intrahippocampal microinfusion of BDNF resulted
in selective upregulation of Arc mRNA and protein, in addition to
rapid and extensive delivery of Arc mRNA transcripts to granule
cell dendrites (Ying et al., 2002).
A recent study has elegantly exploited the fact that Arc and
Homer1a show differential temporal expression patterns in acti-
vated neurons (Guzowski et al., 2001): it was used to mark neu-
ronal ensembles in the GC and Bla that participated in processing
novel taste information. Using in situ hybridization, Saddoris
et al. (2009) showed that repeated exposure to a novel taste
(sucrose) within the time frame required for temporal differen-
tiation of Arc/Homer1a resulted in increased IEG activity, as well
as increased overlap of activated neuronal populations in the GC.
In addition, they showed that odor cues associated with sucrose,
but not with water, elicited potentiation of IEG activity in the GC
similar to that of sucrose itself, independently of Bla. Such cell
populations, responsive to both CS (taste) and US (odor), are held
to be critical for further plasticity.
Another study employed compartmental analysis of temporal
gene transcription fluorescence in situ hybridization (catFISH) for
Arc, relying on complete translocation of Arc from the nucleus to
the cytoplasm over 30 min. This study, which applied Arc catFISH
following a sucrose-conditioned odor preference test involving
nine odor-taste pairings, showed, in contrast to the findings of
Saddoris et al. (2009), that such a flavor experience paradigm
induced a fourfold increase in the percentage of cells activated
by both taste and odor stimulations in the Bla, but not in the IC
(Desgranges et al., 2010). Furthermore, the authors showed that
in odor-conditioned rats the number of cells responsive to one
stimulus was unchanged. An earlier study, which used catFISH
for Arc as well, also found that following CTA, specific Bla neu-
ronal populations were responsive to both CS and US (Barot et al.,
2008). Furthermore, this study demonstrated that when the LI-
CTA paradigm was used, no coincident activation was detected.
The identity of the cells expressing coincident activation remains
to be further characterized.
BRAIN-DERIVED NEUROTROPHIC FACTOR
Brain-derived neurotrophic factor has emerged as a potent medi-
ator both of synaptic plasticity on the cellular level, and of the
interaction of an organism with its environment on the behavioral
level (Moguel-Gonzalez et al., 2008).Along with its tyrosine kinase
receptor TrkB, BDNF plays a critical role in activity-dependent
plasticity processes, such as LTP,learning, and memory (Ma et al.,
2011). Several studies have examined the effect of BDNF in the
CTA paradigm. As mentioned above, BDNF-induced enhance-
ment of CTA retention (Castillo et al., 2006), and this effect
recently has been demonstrated to be dependent on activation
of MAPK and phosphatidylinositol-3-kinase (PI-3K) in the IC
(Castillo and Escobar, 2011). Furthermore, local administration of
BDNF into the IC immediately after similar anisomycin adminis-
tration performed prior to CTA training has been shown to reverse
the anisomycin-induced CTA memory deficits almost to control
levels. Taken together, these results imply that BDNF is a protein-
synthesis-dependent memory enhancer (Moguel-Gonzalez et al.,
2008).
Although BDNF is an IEG, the time scale of its mRNA expres-
sion is somewhat longer than those of Arc or Homer. For instance,
BDNF mRNA expression in the rat IC peaked 6h after CTA, and
began to return to base level after 8 h (Ma et al., 2011). Sur-
prisingly, this study found that BDNF levels in the Bla remained
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 |Volume 5 | Article 87 | 10
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
unchanged, but an increase was observed in the central nuclei of
the amygdala (CeA), peaking at 4 h. The authors showed that phos-
phorylated TrkB levels were elevated in the CeA long before the
CTA-induced BDNF synthesis started, indicating a rapid activity-
dependent BDNF release, presumably independent of protein
synthesis. These data suggest that BDNF secretion and synthe-
sis may be spatially and temporally involved in CTA memory
formation. These authors also demonstrated that BDNF secre-
tion and synthesis were necessary for STM and LTM, respectively,
and, in addition, that BDNF injected into the CeA enhanced the
CTA memory. Furthermore, in another study it was shown that
a human naturally occurring polymorphic variant of BDNF in
knock-in mice (Val66Met) caused impairments in CTA extinc-
tion, but not in its acquisition or retention, whereas in humans,
homozygosity to this variant is associated with altered hippocam-
pal volume, hippocampal-dependent memory impairment, and
susceptibility to neuropsychiatric disorders. On the cellular level,
this polymorphism was associated with alterations in intracellu-
lar trafficking and activity-dependent secretion of Wt BDNF in
neurosecretory cells and cortical neurons (Yu et al., 2009).
FUTURE DIRECTIONS
In all, a more thorough understanding of the molecular mech-
anisms underlying taste memory and of the functioning of the
corresponding regulatory processes in relevant neurons that sub-
serve taste-learning processing circuits should help to elucidate
many important basic aspects of neuronal function. The main
future questions and the directions of future research might be
more general in their nature, reflecting the potential development
of knowledge and understanding in the field of biological mech-
anisms of learning and memory, or they might address questions
specific to the taste systems in the mammalian/human brain.
For example, dissection of the temporal dynamics of any mem-
ory acquisition/consolidation and the establishment of remote
memory retrieval, relearning, and reconsolidation which was
outlined in Figure 2 can be misleading. It is clear that acquisition
and retrieval are well defined in time. However, the consolidation
phase is defined mainly in negative rather than positive terms; it is
sensitive to various disruptions at different time points following
acquisition, and the connections between the various molecu-
lar entities involved in the process of taste learning are not well
described. In Figure 3 we outline some of the processes known to
be taking place in the GC following novel taste learning. This off-
line processing of information in the brain may represent a more
continuous rather than a phasic process. However, the mechanisms
and molecular/cellular participants in these processes are yet to be
identified (Figure 3). It is clear that in order to better understand
the consolidation or off-line processes, it is necessary to enhance
the measurements of the biochemical factors that are correlated
with and necessary for taste learning, and to improve their reso-
lution to a few cubic micrometers or to cellular/subcellular levels.
The tools for this kind of in vivo single-cell analysis have been
dramatically improved recently, and will be used to obtain better
descriptions of the molecular and cellular mechanisms underlying
learning and memorizing processes. Moreover, we and others will
aim to identify the specific circuit or neuronal ensemble involved
in creation and maintenance of any given memory and its value.
Other aims, more specific to the taste system, will be to under-
stand the neuronal mechanisms underlying the taste–memory
condition of waiting on hold” for many hours, with its physi-
cal and chemical information as a conditioned stimulus, pending
arrival of the unconditioned stimulus or the digestive information
that will enable the taste to be tagged as safe or dangerous. This spe-
cific and unique ability of taste-learning beautifully represents the
flexibility of the neuronal system underlying learning and memory
processes, and can also teach us about the limits of the neuronal
systems abilities to create simple and associative memories.
ACKNOWLEDGMENTS
This work was supported by DIP (RO 3971/1-1), ISF (1305/08),
and European Union Seventh Framework Program EUROSPIN
(Contract HEALTH-F2-2009-241498) grants for Kobi Rosenblum.
REFERENCES
Akirav, I., Khatsrinov, V., Vouimba,
R. M., Merhav, M., Fer-
reira, G., Rosenblum, K., and
Maroun, M. (2006). Extinction of
conditioned taste aversion depends
on functional protein synthe-
sis but not on NMDA receptor
activation in the ventromedial
prefrontal cortex. Learn. Mem. 13,
254–258.
Alberini, C. M. (2011). The role
of reconsolidation and the
dynamic process of long-term
memory formation and stor-
age. Front. Behav. Neurosci. 5:2.
doi:10.3389/fnbeh.2011.00012
Alme, M. N., Wibrand, K., Dagestad,
G., and Bramham, C. R. (2007).
Chronic fluoxetine treatment
induces brain region-specific
upregulation of genes associated
with BDNF-induced long-term
potentiation. Neural Plast. 2007,
26496.
Antion, M. D., Merhav, M., Hoeffer, C.
A., Reis, G., Kozma, S. C., Thomas,
G., Schuman, E. M., Rosenblum, K.,
and Klann, E. (2008). Removal of
S6K1 and S6K2 leads to divergent
alterations in learning, memor y, and
synaptic plasticity. Learn. Mem. 15,
29–38.
Banko, J. L., Merhav, M., Stern, E.,
Sonenberg, N., Rosenblum, K., and
Klann, E. (2007). Behavioral alter-
ations in mice lacking the translation
repressor 4E-BP2. Neurobiol. Learn.
Mem. 87, 248–256.
Barki-Harrington, L., Belelovsky, K.,
Doron, G., and Rosenblum, K.
(2009a). “Molecular mechanisms of
taste learning in the insular cor-
tex and amygdala, in Conditioned
Taste Aversion: Behavioral and Neural
Processes, eds S. Reilly and T. R.
Schachtman (Oxford: Oxford Uni-
versity Press), 341–363.
Barki-Harrington, L., Elkobi, A.,
Tzabary, T., and Rosenblum, K.
(2009b). Tyrosine phosphorylation
of the 2B subunit of the NMDA
receptor is necessary for taste
memory formation. J. Neurosci. 29,
9219–9226.
Barot, S. K., Kyono,Y., Clark, E. W., and
Bernstein, I. L. (2008). Visualizing
stimulus convergence in amygdala
neurons during associative learning.
Proc. Natl. Acad. Sci. U.S.A. 105,
20959–20963.
Bartos, J. A., Ulrich, J. D., Li, H.,
Beazely, M. A., Chen,Y., Macdonald,
J. F., and Hell, J. W. (2010). Postsy-
naptic clustering and activation of
Pyk2 by PSD-95. J. Neurosci. 30,
449–463.
Bartsch, D., Ghirardi, M., Skehel, P.
A., Karl, K. A., Herder, S. P., Chen,
M., Bailey, C. H., and Kandel, E.
R. (1995). Aplysia CREB2 represses
long-term facilitation: relief of
repression converts transient facili-
tation into long-term functional and
structural change. Cell 83, 979–992.
Belelovsky, K., Elkobi, A., Kaphzan, H.,
Nairn, A. C., and Rosenblum, K.
(2005). A molecular switch for trans-
lational control in taste memory
consolidation. Eur. J. Neurosci. 22,
2560–2568.
Belelovsky, K., Kaphzan, H., Elkobi, A.,
and Rosenblum, K. (2009). Bipha-
sic activation of the mTOR pathway
in the gustatory cortex is correlated
with and necessary for taste learning.
J. Neurosci. 29, 7424–7431.
Belelovsky, K., Maroun, M., and Rosen-
blum, K. (2007). MAPK activation
in the hippocampus in vivo is cor-
related with experimental setting.
Neurobiol. Learn. Mem. 88, 58–64.
Berman, D. E., Hazvi, S., Neduva, V.,
and Dudai, Y. (2000). The role
of identified neurotransmitter sys-
tems in the response of insular
cortex to unfamiliar taste: activa-
tion of ERK1-2 and formation of
a memory trace. J. Neurosci. 20,
7017–7023.
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 | Volume 5 | Article 87 | 11
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
Berman, D. E., Hazvi, S., Rosen-
blum, K., Seger, R., and Dudai, Y.
(1998). Specific and differential acti-
vation of mitogen-activated pro-
tein kinase cascades by unfamil-
iar taste in the insular cortex of
the behaving rat. J. Neurosci. 18,
10037–10044.
Berman, D. E., Hazvi, S., Stehberg, J.,
Bahar, A., and Dudai, Y. (2003).
Conflicting processes in the extinc-
tion of conditioned taste aversion:
behavioral and molecular aspects
of latency, apparent stagnation, and
spontaneous recovery. Learn. Mem.
10, 16–25.
Besnard, A., Galan-Rodriguez, B.,
Vanhoutte, P., and Caboche, J.
(2011). Elk-1 a transcription
factor with multiple facets in
the brain. Front. Neurosci. 5:35.
doi:10.3389/fnins.2011.00035
Bliss, T. V., and Collingridge, G. L.
(1993). A synaptic model of mem-
ory: long-term potentiation in the
hippocampus. Nature 361, 31–39.
Bramham, C. R. (2007). Control of
synaptic consolidation in the den-
tate gyrus: mechanisms, functions,
and therapeutic implications. Prog.
Brain Res. 163, 453–471.
Bramham, C. R., Alme, M. N., Bittins,
M., Kuipers, S. D., Nair, R. R., Pai,
B., Panja, D., Schubert, M., Soule, J.,
Tiron, A., and Wibrand, K. (2010).
The Arc of synaptic memory. Exp.
Brain Res. 200, 125–140.
Bures, J., Bermudez-Rattoni, F., and
Yamamoto, Y. (1998). Conditioned
Taste Aversion: Memory of a Special
Kind. Oxford, NY:Oxford University
Press.
Burwell, R. D. (2001). Borders and
cytoarchitecture of the perirhinal
and postrhinal cortices in the rat. J.
Comp. Neurol. 437, 17–41.
Cammalleri, M., Lutjens, R., Berton,
F., King, A. R., Simpson, C.,
Francesconi, W., and Sanna, P. P.
(2003). Time-restricted role for den-
dritic activation of the mTOR-
p70S6K pathway in the induction of
late-phase long-term potentiation in
the CA1. Proc. Natl. Acad. Sci. U.S.A.
100, 14368–14373.
Campeau, S., Hayward, M. D., Hope,
B. T., Rosen, J. B., Nestler, E. J., and
Davis, M. (1991). Induction of the c-
fos proto-oncogene in rat amygdala
during unconditioned and condi-
tioned fear. Brain Res. 565, 349–352.
Carleton, A., Accolla, R., and Simon, S.
A. (2010). Coding in the mammalian
gustatory system. Trends Neurosci.
33, 326–334.
Casadio,A., Martin, K. C., Giustetto, M.,
Zhu, H., Chen, M., Bartsch, D., Bai-
ley, C. H., and Kandel, E. R. (1999).
A transient, neuron-wide form of
CREB-mediated long-term facilita-
tion can be stabilized at specific
synapses by local protein synthesis.
Cell 99, 221–237.
Castillo, D. V., and Escobar, M. L.
(2011). A role for MAPK and PI-3K
signaling pathways in brain-derived
neurotrophic factor modification of
conditioned taste aversion retention.
Behav. Brain Res. 217, 248–252.
Castillo, D. V., Figueroa-Guzman,
Y., and Escobar, M. L. (2006).
Brain-derived neurotrophic factor
enhances conditioned taste aversion
retention. Brain Res. 1067, 250–255.
Chen, A., Muzzio, I. A., Malleret, G.,
Bartsch, D., Verbitsky, M., Pavlidis,
P., Yonan, A. L., Vronskaya,S., Grody,
M. B., Cepeda, I., Gilliam, T. C.,
and Kandel, E. R. (2003). Inducible
enhancement of memory storage
and synaptic plasticity in transgenic
mice expressing an inhibitor of ATF4
(CREB-2) and C/EBP proteins. Neu-
ron 39, 655–669.
Chen,A. P.,Ohno, M., Giese, K. P., Kuhn,
R., Chen,R. L., and Silva, A. J. (2006).
Forebrain-specific knockout of B-raf
kinase leads to deficits in hippocam-
pal long-term potentiation, learning,
and memory. J. Neurosci. Res. 83,
28–38.
Chen, X., Nelson, C. D., Li, X., Win-
ters, C. A., Azzam, R., Sousa, A. A.,
Leapman, R. D., Gainer, H., Sheng,
M., and Reese, T. S. (2011). PSD-95
is required to sustain the molecu-
lar organization of the postsynaptic
density. J. Neurosci. 31, 6329–6338.
Chowdhury, S., Shepherd, J. D., Okuno,
H., Lyford, G., Petralia, R. S., Plath,
N., Kuhl,D., Huganir, R. L., and Wor-
ley, P. F. (2006). Arc/Arg3.1 inter-
acts with the endocytic machinery to
regulate AMPA receptor trafficking.
Neuron 52, 445–459.
Cohen-Matsliah, S. I., Brosh, I., Rosen-
blum, K., and Barkai, E. (2007).
A novel role for extracellular
signal-regulated kinase in main-
taining long-term memory-relevant
excitability changes. J. Neurosci. 27,
12584–12589.
Cohen-Matsliah, S. I., Motanis, H.,
Rosenblum, K., and Barkai, E.
(2010). A novel role for protein syn-
thesis in long-term neuronal plastic-
ity: maintaining reduced postburst
afterhyperpolarization. J. Neurosci.
30, 4338–4342.
Collingridge, G. L., Isaac, J. T., and
Wang,Y.T. (2004). Receptortraffick-
ing and synaptic plasticity. Nat. Rev.
Neurosci. 5, 952–962.
Costa-Mattioli, M., Gobert, D.,Stern, E.,
Gamache, K., Colina, R., Cuello, C.,
Sossin, W., Kaufman, R., Pelletier, J.,
Rosenblum,K., Krnjevic,K., Lacaille,
J. C., Nader, K., and Sonenberg, N.
(2007). eIF2alpha phosphorylation
bidirectionally regulates the switch
from short- to long-term synap-
tic plasticity and memory. Cell 129,
195–206.
Costa-Mattioli, M., Sossin, W. S., Klann,
E., and Sonenberg, N. (2009).
Translational control of long-lasting
synaptic plasticity and memory.
Neuron 61, 10–26.
Cousins, S. L., Papadakis, M., Rut-
ter, A. R., and Stephenson, F.
A. (2008). Differential interaction
of NMDA receptor subtypes with
the post-synaptic density-95 family
of membrane associated guanylate
kinase proteins. J. Neurochem. 104,
903–913.
Cui, Z., Lindl, K. A., Mei, B., Zhang, S.,
and Tsien,J. Z. (2005). Requirement
of NMDA receptor reactivation for
consolidation and storage of non-
declarative taste memory revealed
by inducible NR1 knockout. Eur. J.
Neurosci. 22, 755–763.
Davis, H. P., and Squire, L. R. (1984).
Protein synthesis and memory: a
review. Psychol. Bull. 96, 518–559.
de Arce, K. P., Varela-Nallar, L., Farias,
O., Cifuentes, A., Bull, P., Couch, B.
A., Koleske, A. J., Inestrosa, N. C.,
and Alvarez, A. R. (2010). Synaptic
clustering of PSD-95 is regulated by
c-Abl through tyrosine phosphory-
lation. J. Neurosci. 30, 3728–3738.
De la Cruz, V., Rodriguez-Ortiz, C. J.,
Balderas, I., and Bermudez-Rattoni,
F. (2008). Medial temporal lobe
structures participate differentially
in consolidation of safe and aversive
taste memories. Eur. J. Neurosci. 28,
1377–1381.
Delint-Ramirez,I.,Salcedo-Tello, P., and
Bermudez-Rattoni, F. (2008). Spatial
memory formation induces recruit-
ment of NMDA receptor and PSD-
95 to synaptic lipid rafts. J. Neu-
rochem. 106, 1658–1668.
Desgranges, B., Ramirez-Amaya, V.,
Ricano-Cornejo, I., Levy,F., and Fer-
reira, G. (2010). Flavor preference
learning increases olfactory and gus-
tatory convergence onto single neu-
rons in the basolateral amygdala but
not in the insular cortex in rats. PLoS
ONE 5, e10097. doi:10.1371/jour-
nal.pone.0010097
Dong, Y. N., Waxman, E. A., and
Lynch, D. R. (2004). Interactions
of postsynaptic density-95 and the
NMDA receptor 2 subunit con-
trol calpain-mediated cleavage of
the NMDA receptor. J. Neurosci. 24,
11035–11045.
Doron, G., and Rosenblum, K. (2010).
c-Fos expression is elevated in
GABAergic interneurons of the gus-
tatory cortex following novel taste
learning. Neurobiol. Learn. Mem. 94,
21–29.
Dudai, Y. (2004). The neurobiology
of consolidations, or, how stable is
the engram? Annu.Rev.Psychol.55,
51–86.
Elkobi, A., Ehrlich, I., Belelovsky, K.,
Barki-Harrington, L., and Rosen-
blum, K. (2008). ERK-dependent
PSD-95 induction in the gustatory
cortex is necessary for taste learning,
but not retrieval. Nat. Neurosci. 11,
1149–1151.
Escobar, M. L., Alcocer, I., and
Bermudez-Rattoni, F. (2002). In vivo
effects of intracortical administra-
tion of NMDA and metabotropic
glutamate receptors antagonists on
neocortical long-term potentiation
and conditioned taste aversion.
Behav. Brain Res. 129, 101–106.
Escobar, M. L., Alcocer, I., and Chao, V.
(1998). The NMDA receptor antag-
onist CPP impairs conditioned taste
aversion and insular cortex long-
term potentiation in vivo. Brain Res.
812, 246–251.
Escobar, M. L., and Bermudez-Rattoni,
F.(2000). Long-term potentiation in
the insular cortex enhances condi-
tioned taste aversionretention. Brain
Res. 852, 208–212.
Escobar, M. L., Figueroa-Guzman, Y.,
and Gomez-Palacio-Schjetnan, A.
(2003). In vivo insular cortex
LTP induced by brain-derived neu-
rotrophic factor. Brain Res. 991,
274–279.
Fagni, L., Ango, F., Perroy, J., and
Bockaert, J. (2004). Identification
and functional roles of metabotropic
glutamate receptor-interacting pro-
teins. Semin. Cell Dev. Biol. 15,
289–298.
Feldman, D. E. (2009). Synaptic mech-
anisms for plasticity in neocortex.
Annu.Rev.Neurosci.32, 33–55.
Figueroa-Guzman, Y., Kuo, J. S., and
Reilly, S. (2006). NMDA recep-
tor antagonist MK-801 infused
into the insular cortex prevents
the attenuation of gustatory neo-
phobia in rats. Brain Res. 1114,
183–186.
Figueroa-Guzman, Y., and Reilly, S.
(2008). NMDA receptors in the
basolateral amygdala and gusta-
tory neophobia. Brain Res. 1210,
200–203.
Fiore, R. S., Bayer, V. E., Pelech, S. L.,
Posada, J., Cooper, J. A., and Bara-
ban, J. M. (1993). P42 mitogen-
activated protein kinase in brain:
prominent localization in neuronal
cell bodies and dendrites. Neuro-
science 55, 463–472.
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 | Volume 5 | Article 87 | 12
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
Fowler, S. W., Ramsey, A. K., Walker, J.
M., Serfozo, P., Olive, M. F., Schacht-
man, T. R., and Simonyi, A. (2011).
Functional interaction of mGlu5
and NMDA receptors in aversive
learning in rats. Neurobiol. Learn.
Mem. 95, 73–79.
Frank, M. E., Lundy, R. F. Jr., and Con-
treras, R. J. (2008). Cracking taste
codes by tapping into sensory neu-
ron impulse traffic. Prog. Neurobiol.
86, 245–263.
Fujita, S., Adachi, K., Koshikawa, N.,
and Kobayashi, M. (2010). Spa-
tiotemporal dynamics of excitation
in rat insular cortex: intrinsic cor-
ticocortical circuit regulates caudal-
rostro excitatory propagation from
the insular to frontal cortex. Neuro-
science 165, 278–292.
Garcia, J., Kimeldorf,D. J., and Koelling,
R. A. (1955). Conditioned aversion
to saccharin resulting from exposure
to gamma radiation. Science 122,
157–158.
Gibb, S. J., Wolff, M., and Dalrymple-
Alford, J. C. (2006). Odour-place
paired-associate learning and lim-
bic thalamus: comparison of ante-
rior, lateral and medial thalamic
lesions. Behav. Brain Res. 172,
155–168.
Giovannini, M. G. (2006). The role
of the extracellular signal-regulated
kinase pathway in memory encod-
ing. Rev. Neurosci. 17, 619–634.
Gravius, A., Pietraszek, M., Schafer,
D., Schmidt, W. J., and Danysz, W.
(2005). Effects of mGlu1 and mGlu5
receptor antagonists on negatively
reinforced learning. Behav. Pharma-
col. 16, 113–121.
Gutierrez, R., Tellez, L. A., and
Bermudez-Rattoni, F. (2003). Block-
ade of cortical muscarinic but not
NMDA receptors prevents a novel
taste from becoming familiar. Eur. J.
Neurosci. 17, 1556–1562.
Guzman-Ramos, K., Osorio-Gomez, D.,
Moreno-Castilla, P., and Bermudez-
Rattoni, F. (2010). Off-line concomi-
tant release of dopamine and gluta-
mate involvement in taste memory
consolidation. J. Neurochem. 114,
226–236.
Guzowski, J. F., McNaughton, B. L.,
Barnes, C. A., and Worley, P. F.
(2001). Imaging neural activity with
temporal and cellular resolution
using FISH. Curr. Opin. Neurobiol.
11, 579–584.
Guzowski, J. F., Timlin, J. A., Roysam,
B., McNaughton, B. L., Worley, P.
F., and Barnes, C. A. (2005). Map-
ping behaviorally relevant neural
circuits with immediate-early gene
expression. Curr. Opin. Neurobiol.
15, 599–606.
Hay, N., and Sonenberg, N. (2004).
Upstream and downstream of
mTOR. Genes Dev. 18, 1926–1945.
Hinnebusch, J., Cherepanov, P., Du, Y.,
Rudolph, A., Dixon, J. D., Schwan,
T., and Forsberg, A. (2000). Murine
toxin of Yersinia pestis shows phos-
pholipase D activity but is not
required for virulence in mice. Int.
J. Med. Microbiol. 290, 483–487.
Houpt, T. A., and Berlin, R. (1999).
Rapid, labile, and protein synthesis-
independent short-term memory in
conditioned taste aversion. Learn.
Mem. 6, 37–46.
Izquierdo, I., Bevilaqua, L. R., Rossato,
J. I., Bonini, J. S., Da Silva, W. C.,
Medina, J. H., and Cammarota, M.
(2006). The connection between the
hippocampal and the striatal mem-
ory systems of the brain: a review
of recent findings. Neurotox. Res. 10,
113–121.
Jones, M. W., French, P. J., Bliss, T. V.,
and Rosenblum, K. (1999). Molecu-
lar mechanisms of long-term poten-
tiation in the insular cortex in vivo.
J. Neurosci. 19, RC36.
Kalia, L. V., Pitcher, G. M., Pelkey, K. A.,
and Salter, M. W. (2006). PSD-95 is
a negative regulator of the tyrosine
kinase Src in the NMDA receptor
complex. EMBO J. 25, 4971–4982.
Kandel, E. R. (2001). The molecular
biology of memory storage: a dia-
logue between genes and synapses.
Science 294, 1030–1038.
Kaphzan, H., Doron, G., and Rosen-
blum, K. (2007). Co-application
of NMDA and dopamine-induced
rapid translation of RSK2 in the
mature hippocampus. J. Neurochem.
103, 388–399.
Kaphzan, H., ORiordan, K. J., Man-
gan, K. P., Levenson, J. M., and
Rosenblum, K. (2006). NMDA and
dopamine converge on the NMDA-
receptor to induce ERK activation
and synaptic depression in mature
hippocampus. PLoS ONE 1, e138.
doi:10.1371/journal.pone.0000138
Kaufmann, W. E., and Worley, P. F.
(1999). Neural activity and immedi-
ate early gene expression in the cere-
bral cortex. Dev. Disabil. Res. Rev. 5,
41–50.
Koh, M. T., and Bernstein, I. L. (2005).
Mapping conditioned taste aversion
associations using c-Fos reveals a
dynamic role for insular cortex.
Behav. Neurosci. 119, 388–398.
Kornau, H. C., Schenker, L. T.,
Kennedy, M. B., and Seeburg,
P. H. (1995). Domain interac-
tion between NMDA receptor sub-
units and the postsynaptic den-
sity protein PSD-95. Science 269,
1737–1740.
Koya, E., Golden, S. A., Harvey, B. K.,
Guez-Barber,D. H., Berkow,A., Sim-
mons, D. E., Bossert, J. M., Nair,
S. G., Uejima, J. L., Marin, M. T.,
Mitchell, T. B., Farquhar, D., Ghosh,
S. C., Mattson, B. J., and Hope,
B. T. (2009). Targeted disruption
of cocaine-activated nucleus accum-
bens neurons prevents context-
specific sensitization. Nat. Neurosci.
12, 1069–1073.
Lamprecht, R., and Dudai, Y. (1996).
Transient expression of c-Fos in rat
amygdala during training is required
for encoding conditioned taste aver-
sion memory. Learn. Mem. 3, 31–41.
Lavezzari, G., McCallum, J., Lee, R.,
and Roche, K. W. (2003). Differ-
ential binding of the AP-2 adap-
tor complex and PSD-95 to the
C-terminus of the NMDA recep-
tor subunit NR2B regulates surface
expression. Neuropharmacology 45,
729–737.
Li, S., Gu, Y., Meng, B., Mei, B.,and Li, F.
(2010). The different effects of over-
expressing murine NMDA receptor
2B subunit in the forebrain on con-
ditioned taste aversion. Brain Res.
1351, 165–171.
Lin, P. Y., Wang, S. P., Tai, M. Y.,
and Tsai, Y. F. (2010). Differen-
tial involvement of medial pre-
frontal cortex and basolateral amyg-
dala extracellular signal-regulated
kinase in extinction of conditioned
taste aversion is dependent on dif-
ferent intervals of extinction follow-
ing conditioning. Neuroscience 171,
125–133.
Lindquist, D. H.,Vogel, R. W., and Stein-
metz, J. E. (2009). Associative and
non-associative blinking in classi-
cally conditioned adult rats. Physiol.
Behav. 96, 399–411.
Liraz, O., Rosenblum, K., and
Barkai, E. (2009). CAMKII
activation is not required
for maintenance of learning-
induced enhancement of neuronal
excitability. PLoS ONE 4, e4289.
doi:10.1371/journal.pone.0004289
Loewith, R., Jacinto, E., Wullschleger, S.,
Lorberg, A., Crespo, J. L., Bonen-
fant, D., Oppliger, W., Jenoe, P.,
and Hall, M. N. (2002). Two TOR
complexes, only one of which is
rapamycin sensitive, have distinct
roles in cell growth control. Mol. Cell
10, 457–468.
Lu, P. D., Harding, H. P., and Ron,
D. (2004). Translation reinitiation at
alternative open reading frames reg-
ulates gene expression in an inte-
grated stress response. J. Cell Biol.
167, 27–33.
Lubow, R. E. (1989). Latent Inhibition
and Conditioned Attention Theory.
Cambridge: Cambridge University
Press.
Lyford, G. L., Yamagata, K., Kaufmann,
W. E., Barnes, C. A., Sanders, L. K.,
Copeland, N. G., Gilbert, D. J., Jenk-
ins, N. A., Lanahan, A. A., and Wor-
ley, P. F. (1995). Arc,a growth factor
and activity-regulated gene, encodes
a novel cytoskeleton-associated pro-
tein that is enriched in neuronal
dendrites. Neuron 14, 433–445.
Ma, L., Wang, D. D., Zhang, T. Y.,Yu, H.,
Wang,Y.,Huang, S. H., Lee, F. S., and
Chen, Z. Y. (2011). Region-specific
involvement of BDNF secretion and
synthesis in conditioned taste aver-
sion memory formation. J. Neurosci.
31, 2079–2090.
Malenka, R. C., and Bear, M. F. (2004).
LTP and LTD: an embarrassment of
riches. Neuron 44, 5–21.
Mao, L., and Wang, J. Q. (2002).
Interactions between ionotropic and
metabotropic glutamate receptors
regulate cAMP response element-
binding protein phosphorylation in
cultured striatal neurons. Neuro-
science 115, 395–402.
Mattson, B. J., Koya, E., Simmons,
D. E., Mitchell, T. B., Berkow, A.,
Crombag, H. S., and Hope, B. T.
(2008). Context-specific sensitiza-
tion of cocaine-induced locomo-
tor activity and associated neuronal
ensembles in rat nucleus accumbens.
Eur. J. Neurosci. 27, 202–212.
Mazzucchelli, C., Vantaggiato, C.,
Ciamei, A., Fasano, S., Pakhotin,
P., Krezel, W., Welzl, H., Wolfer,
D. P., Pages, G., Valverde, O.,
Marowsky, A., Porrazzo, A., Orban,
P. C., Maldonado, R., Ehrengruber,
M. U., Cestari, V., Lipp, H. P.,
Chapman, P. F., Pouyssegur, J., and
Brambilla, R. (2002). Knockout
of ERK1 MAP kinase enhances
synaptic plasticity in the striatum
and facilitates striatal-mediated
learning and memory. Neuron 34,
807–820.
McGaugh, J.L. (2000). Memory a cen-
tury of consolidation. Science 287,
248–251.
McGeachie, A. B., Cingolani, L. A.,
and Goda, Y. (2011). A stabilising
influence: integrins in regulation of
synaptic plasticity. Neurosci. Res. 70,
24–29.
Mei, F., Nagappan, G., Ke, Y., Sack-
tor, T. C., and Lu, B. (2011). BDNF
facilitates L-LTP maintenance in the
absence of protein synthesis through
PKMzeta. PLoS ONE 6, e21568.
doi:10.1371/journal.pone.0021568
Meiri, N., and Rosenblum, K. (1998).
Lateral ventricle injection of the pro-
tein synthesis inhibitor anisomycin
impairs long-term memory in a
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 | Volume 5 | Article 87 | 13
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
spatial memory task. Brain Res. 789,
48–55.
Merhav, M., Kuulmann-Vander, S.,
Elkobi, A., Jacobson-Pick, S.,
Karni, A., and Rosenblum, K.
(2006). Behavioral interference
and C/EBPbeta expression in the
insular-cortex reveal a prolonged
time period for taste memory
consolidation. Learn. Mem. 13,
571–574.
Merhav,M., and Rosenblum, K. (2008).
Facilitation of taste memory acquisi-
tion by experiencing previous novel
taste is protein-synthesis dependent.
Learn. Mem. 15, 501–507.
Messaoudi, E., Kanhema, T., Soule, J.,
Tiron, A., Dagyte, G., da Silva, B.,
and Bramham, C. R. (2007). Sus-
tained Arc/Arg3.1 synthesis controls
long-term potentiation consolida-
tion through regulation of local
actin polymerization in the den-
tate gyrus in vivo. J. Neurosci. 27,
10445–10455.
Miranda, M. I., and Bermudez-Rattoni,
F. (1999). Reversible inactivation of
the nucleus basalis magnocellularis
induces disruption of cortical acetyl-
choline release and acquisition, but
not retrieval, of aversive memo-
ries. Proc. Natl. Acad. Sci. U.S.A. 96,
6478–6482.
Miranda, M. I., and Bermudez-Rattoni,
F. (2007). Cholinergic activity in the
insular cortex is necessary for acqui-
sition and consolidation of con-
textual memory. Neurobiol. Learn.
Mem. 87, 343–351.
Miranda, M. I., Ramirez-Lugo, L., and
Bermudez-Rattoni, F. (2000). Corti-
cal cholinergic activity is related to
the novelty of the stimulus. Brain
Res. 882, 230–235.
Miranda, M. I., Rodriguez-Garcia, G.,
Reyes-Lopez, J. V., Ferr y, B., and Fer-
reira, G. (2008). Differential effects
of beta-adrenergic receptor block-
ade in basolateral amygdala or insu-
lar cortex on incidental and associa-
tive taste learning. Neurobiol. Learn.
Mem. 90, 54–61.
Moguel-Gonzalez, M., Gomez-Palacio-
Schjetnan, A., and Escobar, M. L.
(2008). BDNF reverses the CTA
memory deficits produced by inhibi-
tion of protein synthesis. Neurobiol.
Learn. Mem. 90, 584–587.
Morris, R., Frey, S., Kasambira, T., and
Petrides, M. (1999). Ibotenic acid
lesions of the basolateral, but not
the central, amygdala interfere with
conditioned taste aversion: evidence
from a combined behavioral and
anatomical tract-tracing investiga-
tion. Behav. Neurosci. 113, 291–302.
Morris, R. G. (2006). Elements of a neu-
robiological theory of hippocampal
function: the role of synaptic plas-
ticity, synaptic tagging and schemas.
Eur. J. Neurosci. 23, 2829–2846.
Nachman, M., and Ashe, J. H. (1974).
Effects of basolateral amygdala
lesions on neophobia, learned taste
aversions, and sodium appetite in
rats. J. Comp. Physiol. Psychol. 87,
622–643.
Nairn, A. C., and Palfrey, H. C. (1987).
Identification of the major Mr
100,000 substrate for calmodulin-
dependent protein kinase III
in mammalian cells as elonga-
tion factor-2. J. Biol. Chem. 262,
17299–17303.
Naisbitt, S., Kim, E., Tu, J. C., Xiao,
B., Sala, C., Valtschanoff, J., Wein-
berg, R. J., Worley, P. F., and Sheng,
M. (1999). Shank, a novel family
of postsynaptic density proteins that
binds to the NMDA receptor/PSD-
95/GKAP complex and cortactin.
Neuron 23, 569–582.
Navakkode, S., Sajikumar, S., Sack-
tor, T. C., and Frey, J. U. (2010).
Protein kinase Mzeta is essential
for the induction and mainte-
nance of dopamine-induced long-
term potentiation in apical CA1 den-
drites. Learn. Mem. 17, 605–611.
Neves, G., Cooke, S. F., and Bliss, T. V.
(2008). Synaptic plasticity, memory
and the hippocampus: a neural net-
work approach to causality.Nat. Rev.
Neurosci. 9, 65–75.
Nunez-Jaramillo, L., Jimenez, B.,
Ramirez-Munguia, N., Delint-
Ramirez, I., Luna-Illades, C., Tapia,
R., and Bermudez-Rattoni, F.
(2008). Taste novelty induces intra-
cellular redistribution of NR2A and
NR2B subunits of NMDA receptor
in the insular cortex. Brain Res.
1215, 116–122.
Nunez-Jaramillo, L., Ramirez-Lugo, L.,
Herrera-Morales, W., and Miranda,
M. I. (2010). Taste memory forma-
tion: latest advances and challenges.
Behav. Brain Res. 207, 232–248.
Panja, D., Dagyte, G., Bidinosti, M.,
Wibrand, K., Kristiansen, A. M.,
Sonenberg, N., and Bramham, C. R.
(2009). Novel translational control
in Arc-dependent long term poten-
tiation consolidation in vivo. J. Biol.
Chem. 284, 31498–31511.
Parvez, S., Ramachandran, B., and Frey,
J. U. (2010). Functional differences
between and across different regions
of the apical branch of hippocam-
pal CA1 dendrites with respect to
long-term depression induction and
synaptic cross-tagging. J. Neurosci.
30, 5118–5123.
Pastalkova, E., Serrano, P., Pinkhasova,
D., Wallace, E., Fenton, A. A., and
Sacktor, T. C. (2006). Storage of
spatial information by the mainte-
nance mechanism of LTP. Science
313, 1141–1144.
Peebles, C. L., Yoo, J., Thwin, M. T.,
Palop, J. J., Noebels, J. L., and
Finkbeiner, S. (2010). Arc regulates
spine morphology and maintains
network stability in vivo. Proc. Natl.
Acad. Sci. U.S.A. 107, 18173–18178.
Plath, N., Ohana, O., Dammermann,
B., Errington, M. L., Schmitz, D.,
Gross, C., Mao, X., Engelsberg, A.,
Mahlke, C., Welzl, H., Kobalz, U.,
Stawrakakis, A., Fernandez, E., Wal-
tereit, R., Bick-Sander, A., Therstap-
pen, E., Cooke, S. F., Blanquet, V.,
Wurst, W., Salmen, B., Bosl, M. R.,
Lipp, H. P., Grant, S. G., Bliss, T. V.,
Wolfer, D. P., and Kuhl, D. (2006).
Arc/Arg3.1 is essential for the con-
solidation of synaptic plasticity and
memories. Neuron 52, 437–444.
Pozo, K., and Goda, Y. (2010). Unrav-
eling mechanisms of homeosta-
tic synaptic plasticity. Neuron 66,
337–351.
Proud, C. G. (2000). “Control of the
elongation phase of protein synthe-
sis, in Translational Control of Gene
Expression, eds N. Sonenberg, J. Her-
shy, and M. Mathews (Cold Spring
Harbor, NY: Cold Spring Harbor
Laboratory Press), 719–739.
Ramirez-Lugo, L., Miranda, M. I.,
Escobar, M. L., Espinosa, E., and
Bermudez-Rattoni, F. (2003). The
role of cortical cholinergic pre-
and post-synaptic receptors in
taste memory formation. Neurobiol.
Learn. Mem. 79, 184–193.
Raught, B., Gingras, A. C., and
Sonenberg, N. (2001). The tar-
get of rapamycin (TOR) proteins.
Proc. Natl. Acad. Sci. U.S.A. 98,
7037–7044.
Rodriguez-Duran, L. F., Castillo, D.
V., Moguel-Gonzalez, M., and Esco-
bar, M. L. (2011). Conditioned taste
aversion modifies persistently the
subsequent induction of neocorti-
cal long-term potentiation in vivo.
Neurobiol. Learn. Mem. 95, 519–526.
Roman, C., Nebieridze, N., Sastre, A.,
and Reilly, S. (2006). Effects of
lesions of the bed nucleus of the
stria terminalis, lateral hypothala-
mus, or insular cortex on condi-
tioned taste aversion and condi-
tioned odor aversion. Behav. Neu-
rosci. 120, 1257–1267.
Roman, C., and Reilly, S. (2007). Effects
of insular cortex lesions on condi-
tioned taste aversion and latent inhi-
bition in the rat. Eur. J. Neurosci. 26,
2627–2632.
Rondard, P., Goudet, C., Kniazeff, J.,
Pin, J. P., and Prezeau, L. (2011).
The complexity of their activation
mechanism opens new possibilities
for the modulation of mGlu and
GABAB class C G protein-coupled
receptors. Neuropharmacology 60,
82–92.
Rong, Y., Lu, X., Bernard, A.,
Khrestchatisky, M., and Baudry, M.
(2001). Tyrosine phosphorylation
of ionotropic glutamate receptors by
Fyn or Src differentially modulates
their susceptibility to calpain and
enhances their binding to spectrin
and PSD-95. J. Neurochem. 79,
382–390.
Rosenblum, K. (2008). “Conditioned
taste aversion and taste learning:
molecular mechanisms,” in Concise
Learning and Memory: The Editors
Selection, ed. J. H. Byrne (Ams-
terdam: Academic Press, Elsevier),
465–479.
Rosenblum, K., Berman, D. E., Hazvi, S.,
Lamprecht, R., and Dudai, Y. (1997).
NMDA receptor and the tyrosine
phosphorylation of its 2B subunit
in taste learning in the rat insular
cortex. J. Neurosci. 17, 5129–5135.
Rosenblum, K., Dudai, Y., and Richter-
Levin, G. (1996). Long-term
potentiation increases tyrosine
phosphorylation of the N-methyl-
D-aspartate receptor subunit
2B in rat dentate gyrus in vivo.
Proc. Natl. Acad. Sci. U.S.A. 93,
10457–10460.
Rosenblum, K., Futter, M., Voss, K.,
Erent, M., Skehel, P. A., French, P.,
Obosi, L., Jones, M. W., and Bliss, T.
V. (2002). The role of extracellular
regulated kinases I/II in late-phase
long-term potentiation. J. Neurosci.
22, 5432–5441.
Rosenblum, K., Meiri, N., and Dudai,
Y. (1993). Taste memory: the role of
protein synthesis in gustatory cortex.
Behav. Neural Biol. 59, 49–56.
Rosenblum, K., Schul, R., Meiri, N.,
Hadari, Y. R., Zick, Y., and Dudai, Y.
(1995). Modulation of protein tyro-
sine phosphorylation in rat insular
cortex after conditioned taste aver-
sion training. Proc. Natl. Acad. Sci.
U.S.A. 92, 1157–1161.
Ryazanov, A. G., and Davydova, E.
K. (1989). Mechanism of elon-
gation factor 2 (EF-2) inactiva-
tion upon phosphorylation. Phos-
phorylated EF-2 is unable to cat-
alyze translocation. FEBS Lett. 251,
187–190.
Sacktor, T. C. (2008). PKMzeta, LTP
maintenance, and the dynamic mol-
ecular biology of memory storage.
Prog. Brain Res. 169, 27–40.
Saddoris, M. P., Holland, P. C., and
Gallagher, M. (2009). Associatively
learned representations of taste
outcomes activate taste-encoding
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 | Volume 5 | Article 87 | 14
Gal-Ben-Ari and Rosenblum Molecular mechanisms of taste memory
neural ensembles in gustatory cor-
tex. J. Neurosci. 29, 15386–15396.
Sajikumar, S., and Korte, M. (2011).
Different compartments of apical
CA1 dendrites have different plastic-
ity thresholds for expressing synaptic
tagging and capture. Learn. Mem. 18,
327–331.
Schachtman, T. R., Bills, C., Ghinescu,
R., Murch, K., Serfozo, P., and
Simonyi, A. (2003). MPEP, a selec-
tive metabotropic glutamate recep-
tor 5 antagonist, attenuates condi-
tioned taste aversion in rats. Behav.
Brain Res. 141, 177–182.
Seger, R., and Krebs, E. G. (1995). The
MAPK signaling cascade. FASEB J. 9,
726–735.
Seroussi, Y., Brosh, I., and Barkai,
E. (2002). Learning-induced
reduction in post-burst after-
hyperpolarization (AHP) is
mediated by activation of PKC.
Eur. J. Neurosci. 16, 965–969.
Shema, R., Haramati, S., Ron, S., Hazvi,
S., Chen, A., Sacktor, T. C., and
Dudai, Y. (2011). Enhancement of
consolidated long-term memory by
overexpression of protein kinase
Mzeta in the neocortex. Science 331,
1207–1210.
Shema, R., Hazvi, S., Sacktor, T. C., and
Dudai, Y. (2009). Boundary condi-
tions for the maintenance of mem-
ory by PKMzeta in neocortex. Learn.
Mem. 16, 122–128.
Shepherd, J. D., and Bear, M. F. (2011).
New views of Arc, a master regulator
of synaptic plasticity. Nat. Neurosci.
14, 279–284.
Shepherd, J. D., Rumbaugh, G., Wu,
J., Chowdhury, S., Plath, N., Kuhl,
D., Huganir, R. L., and Worley, P. F.
(2006). Arc/Arg3.1 mediates home-
ostatic synaptic scaling of AMPA
receptors. Neuron 52, 475–484.
Simonyi, A., Serfozo, P., Parker, K. E.,
Ramsey, A. K., and Schachtman,
T. R. (2009). Metabotropic gluta-
mate receptor 5 in conditioned taste
aversion learning. Neurobiol. Learn.
Mem. 92, 460–463.
Simonyi, A., Serfozo, P., Shelat, P. B.,
Dopheide, M. M., Coulibaly, A.
P., and Schachtman, T. R. (2007).
Differential roles of hippocampal
metabotropic glutamate receptors
1 and 5 in inhibitory avoidance
learning. Neurobiol. Learn. Mem. 88,
305–311.
Sjostrom, P. J., Rancz, E. A., Roth, A.,
and Hausser, M. (2008). Dendritic
excitability and synaptic plasticity.
Physiol. Rev. 88, 769–840.
St Andre, J., and Reilly,S. (2007). Effects
of central and basolateral amygdala
lesions on conditioned taste aversion
and latent inhibition. Behav. Neu-
rosci. 121, 90–99.
Swank, M. W., and Sweatt, J. D. (2001).
Increased histone acetyltransferase
and lysine acetyltransferase activ-
ity and biphasic activation of the
ERK/RSK cascade in insular cor-
tex during novel taste learning. J.
Neurosci. 21, 3383–3391.
Sweatt, J. D. (2001). The neuronal MAP
kinase cascade: a biochemical sig-
nal integration system subserving
synaptic plasticity and memory. J.
Neurochem. 76, 1–10.
Sweetat, S., Rosenblum, K., and Lam-
precht, R. (2011). Rho-associated
kinase in the gustatory cortex is
involved in conditioned taste aver-
sion memory formation but not
in memory retrieval or relearning.
Neurobiol. Learn. Mem.
Tang, S. J., Reis, G., Kang, H., Gingras,
A. C., Sonenberg, N., and Schuman,
E. M. (2002). A rapamycin-sensitive
signaling pathway contributes to
long-term synaptic plasticity in the
hippocampus. Proc. Natl. Acad. Sci.
U.S.A. 99, 467–472.
Tirosh, S., Elkobi, A., Rosenblum, K.,
and Meiri, N. (2007). A role for
eukaryotic translation initiation fac-
tor 2B (eIF2B) in taste mem-
ory consolidation and in ther-
mal control establishment during
the critical period for sensory
development. Dev. Neurobiol. 67,
728–739.
Tsokas, P., Ma, T., Iyengar, R., Landau,
E. M., and Blitzer, R. D. (2007).
Mitogen-activated protein kinase
upregulates the dendritic translation
machinery in long-term potentia-
tion by controlling the mammalian
target of rapamycin pathway. J. Neu-
rosci. 27, 5885–5894.
Tsui, C. C., Copeland, N. G., Gilbert,
D. J., Jenkins, N. A., Barnes, C., and
Worley, P. F. (1996). Narp, a novel
member of the pentraxin family,
promotes neurite outgrowth and is
dynamically regulated by neuronal
activity. J. Neurosci. 16, 2463–2478.
Tu, J. C., Xiao, B., Naisbitt, S., Yuan,
J. P., Petralia, R. S., Brakeman,
P., Doan, A., Aakalu, V. K., Lana-
han, A. A., Sheng, M., and Wor-
ley, P. F. (1999). Coupling of
mGluR/Homer and PSD-95 com-
plexes by the Shank family of postsy-
naptic density proteins. Neuron 23,
583–592.
Vantaggiato, C., Formentini, I., Bon-
danza, A., Bonini, C., Naldini, L.,
and Brambilla, R. (2006). ERK1 and
ERK2 mitogen-activated protein
kinases affect Ras-dependent cell
signaling differentially. J. Biol. 5, 14.
Vattem, K. M., and Wek, R. C. (2004).
Reinitiation involving upstream
ORFs regulates ATF4 mRNA
translation in mammalian cells.
Proc. Natl. Acad. Sci. U.S.A. 101,
11269–11274.
Vlachos, A., Maggio, N., and Jedlicka, P.
(2008). Just in time for late-LTP: a
mechanism for the role of PKMzeta
in long-term memory. Commun.
Integr. Biol. 1, 190–191.
von Kraus, L. M., Sacktor, T. C., and
Francis, J. T. (2010). Erasing sen-
sorimotor memories via PKMzeta
inhibition. PLoS ONE 5, e11125.
doi:10.1371/journal.pone.0011125
Westmark, P. R., Westmark, C. J., Wang,
S., Levenson, J., ORiordan, K. J.,
Burger, C., and Malter, J. S. (2010).
Pin1 and PKMzeta sequentially con-
trol dendritic protein synthesis. Sci.
Signal. 3, ra18.
Xiao, B., Tu, J. C., Petralia, R. S.,
Yuan, J. P., Doan, A., Breder, C.
D., Ruggiero, A., Lanahan, A. A.,
Wenthold, R. J., and Worley, P.
F. (1998). Homer regulates the
association of group 1 metabotropic
glutamate receptors with multiva-
lent complexes of homer-related,
synaptic proteins. Neuron 21,
707–716.
Yamamoto, K., Koyanagi,Y., Koshikawa,
N., and Kobayashi, M. (2010).
Postsynaptic cell type-dependent
cholinergic regulation of GABAer-
gic synaptic transmission in rat
insular cortex. J. Neurophysiol. 104,
1933–1945.
Yamamoto, T., Shimura, T., Sako, N.,
Yasoshima, Y., and Sakai, N. (1994).
Neural substrates for conditioned
taste aversion in the rat. Behav.Brain
Res. 65, 123–137.
Yamamoto, T., Yuyama, N., Kato, T.,
and Kawamura, Y. (1984). Gusta-
tory responses of cortical neurons
in rats. I. Response characteristics. J.
Neurophysiol. 51, 616–635.
Yamamoto, T., Yuyama, N., Kato, T.,
and Kawamura, Y. (1985). Gusta-
tory responses of cortical neurons in
rats. III. Neural and behavioral mea-
sures compared. J. Neurophysiol. 53,
1370–1386.
Yang, L., Mao, L., Tang, Q., Sam-
dani, S., Liu, Z., and Wang, J. Q.
(2004). A novel Ca2+-independent
signaling pathway to extracellular
signal-regulated protein kinase by
coactivation of NMDA receptors
and metabotropic glutamate recep-
tor5inneurons.J. Neurosci. 24,
10846–10857.
Yasoshima, Y., Morimoto, T., and
Yamamoto, T. (2000). Different dis-
ruptive effects on the acquisition
and expression of conditioned taste
aversion by blockades of amygdalar
ionotropic and metabotropic gluta-
matergic receptor subtypes in rats.
Brain Res. 869, 15–24.
Yefet, K., Merhav, M., Kuulmann-
Vander, S., Elkobi, A., Belelovsky,
K., Jacobson-Pick, S., Meiri, N., and
Rosenblum, K. (2006). Different sig-
nal transduction cascades are acti-
vated simultaneously in the rat insu-
lar cortex and hippocampus fol-
lowing novel taste learning. Eur. J.
Neurosci. 24, 1434–1442.
Ying, S. W., Futter, M., Rosenblum, K.,
Webber, M. J., Hunt, S. P., Bliss,
T. V., and Bramham, C. R. (2002).
Brain-derived neurotrophic factor
induces long-term potentiation in
intact adult hippocampus: require-
ment for ERK activation coupled
to CREB and upregulation of Arc
synthesis. J. Neurosci. 22, 1532–1540.
Yu, H., Wang, Y., Pattwell, S., Jing,
D., Liu, T., Zhang, Y., Bath, K. G.,
Lee, F. S., and Chen, Z. Y. (2009).
Variant BDNF Val66Met polymor-
phism affects extinction of condi-
tioned aversive memory. J. Neurosci.
29, 4056–4064.
Zalewska, T., Ziemka-Nalecz, M., and
Domanska-Janik, K. (2005). Tran-
sient forebrain ischemia effects
interaction of Src, FAK, and PYK2
with the NR2B subunit of N-
methyl-D-aspartate receptor in ger-
bil hippocampus. Brain Res. 1042,
214–223.
Conflict of Interest Statement: The
authors declare that the research was
conducted in the absence of any
commercial or financial relationships
that could be construed as a potential
conflict of interest.
Received: 29 August 2011; accepted: 12
December 2011; published online: 05 Jan-
uary 2012.
Citation: Gal-Ben-Ari S and Rosen-
blum K (2012) Molecular mecha-
nisms underlying memory consolida-
tion of taste information in the cor-
tex. Front. Behav. Neurosci. 5:87. doi:
10.3389/fnbeh.2011.00087
Copyright © 2012 Gal-Ben-Ari and
Rosenblum. This is an open-access arti-
cle distributed under the terms of
the Creative Commons Attribution Non
Commercial License, which permits non-
commercial use, distribution, and repro-
duction in other forums, provided the
original authors and source are credited.
Frontiers in Behavioral Neuroscience www.frontiersin.org January 2012 | Volume 5 | Article 87 | 15
... While previous work on CTA has focused on biochemical signals [83][84][85] , such as ERK [86][87][88]34,89,90 , here we have taken a complementary approach that emphasizes the role of neural activity across the stages of learning. Of course, biochemistry and neural activity function together and interact with each other, and thus future work will focus on bridging these levels of understanding. ...
Preprint
Full-text available
Animals learn the value of foods based on their postingestive effects and thereby develop aversions to foods that are toxic and preferences to those that are nutritious. However, it remains unclear how the brain is able to assign credit to flavors experienced during a meal with postingestive feedback signals that can arise after a substantial delay. Here, we reveal an unexpected role for postingestive reactivation of neural flavor representations in this temporal credit assignment process. To begin, we leverage the fact that mice learn to associate novel, but not familiar, flavors with delayed gastric malaise signals to investigate how the brain represents flavors that support aversive postingestive learning. Surveying cellular resolution brainwide activation patterns reveals that a network of amygdala regions is unique in being preferentially activated by novel flavors across every stage of the learning process: the initial meal, delayed malaise, and memory retrieval. By combining high-density recordings in the amygdala with optogenetic stimulation of genetically defined hindbrain malaise cells, we find that postingestive malaise signals potently and specifically reactivate amygdalar novel flavor representations from a recent meal. The degree of malaise-driven reactivation of individual neurons predicts strengthening of flavor responses upon memory retrieval, leading to stabilization of the population-level representation of the recently consumed flavor. In contrast, meals without postingestive consequences degrade neural flavor representations as flavors become familiar and safe. Thus, our findings demonstrate that interoceptive reactivation of amygdalar flavor representations provides a neural mechanism to resolve the temporal credit assignment problem inherent to postingestive learning.
... The IC is an integrating forebrain structure involved in sensory perception, learning, and memory [31]. It receives afferent projections from thalamic nuclei and forms reciprocal connections with the amygdala, ACC, and other cortical association areas [32,33]. ...
Article
Full-text available
Cumulative animal and human studies have consistently demonstrated that two major cortical regions in the brain, namely the anterior cingulate cortex (ACC) and insular cortex (IC), play critical roles in pain perception and chronic pain. Neuronal synapses in these cortical regions of adult animals are highly plastic and can undergo long-term potentiation (LTP), a phenomenon that is also reported in brain areas for learning and memory (such as the hippocampus). Genetic and pharmacological studies show that inhibiting such cortical LTP can help to reduce behavioral sensitization caused by injury as well as injury-induced emotional changes. In this review, we will summarize recent progress related to synaptic mechanisms for different forms of cortical LTP and their possible contribution to behavioral pain and emotional changes.
... The gustatory cortex (GC, which processes taste [70,71,75,129]) and visceral insular cortex (VIC, which processes visceral stimuli [75,76,130,131]) are topographically distributed throughout the insular cortex, including layers IV, V, and VI [75,76,130,132,133]. Our data demonstrate that the GC and VIC show significant overlap in Fos expression [75,134]. The GC reacts to novel taste and lesions are known to cause a decrease in neophobia [58,71,129,135]. ...
Article
Full-text available
Females are more affected by psychiatric illnesses including eating disorders, depression, and post-traumatic stress disorder than males. However, the neural mechanisms mediating these sex differences are poorly understood. Animal models can be useful in exploring such neural mechanisms. Conditioned taste aversion (CTA) is a behavioral task that assesses how animals process the competition between associated reinforcing and aversive stimuli in subsequent task performance, a process critical to healthy behavior in many domains. The purpose of the present study was to identify sex differences in this behavior and associated neural responses. We hypothesized that females would value the rewarding stimulus (Boost®) relative to the aversive stimulus (LiCl) more than males in performing CTA. We evaluated behavior (Boost® intake, LiCl-induced behaviors, ultrasonic vocalizations (USVs), CTA performance) and Fos activation in relevant brain regions after the acute stimuli [acute Boost® (AB), acute LiCl (AL)] and the context-only task control (COT), Boost® only task (BOT) and Boost®-LiCl task (BLT). Acutely, females drank more Boost® than males but showed similar aversive behaviors after LiCl. Females and males performed CTA similarly. Both sexes produced 55 kHz USVs anticipating BOT and inhibited these calls in the BLT. However, more females emitted both 22 kHz and 55 kHz USVs in the BLT than males: the latter correlated with less CTA. Estrous cycle stage also influenced 55 kHz USVs. Fos responses were similar in males and females after AB or AL. Females engaged the gustatory cortex and ventral tegmental area (VTA) more than males during the BOT and males engaged the amygdala more than females in both the BOT and BLT. Network analysis of correlated Fos responses across brain regions identified two unique networks characterizing the BOT and BLT, in both of which the VTA played a central role. In situ hybridization with RNAscope identified a population of D1-receptor expressing cells in the CeA that responded to Boost® and D2 receptor-expressing cells that responded to LiCl. The present study suggests that males and females differentially process the affective valence of a stimulus to produce the same goal-directed behavior.
... In the insular cortex, glutamate plays a critical role in the regulation of appetitive and aversive taste memory (Bermúdez-Rattoni, 2004;Gal-Ben-Ari. and Rosenblum, 2012;Núñez-Jaramillo et al., 2010). In particular, regarding appetitive memory, the NMDA receptor antagonist MK-801 in the NAc was able to inhibit the reconsolidation of instrumental memory for sucrose (Piva et al., 2018). Moreover, ketamine, another NMDA receptor antagonist, given 24 h prior to renewal of sucrose-seeking, significantly inhib ...
Article
Dopaminergic neurons projecting from the Substantia Nigra to the Striatum play a critical role in motor functions while dopaminergic neurons originating in the Ventral Tegmental Area (VTA) and projecting to the Nucleus Accumbens, Hippocampus and other cortical structures regulate rewarding learning. While VTA mainly consists of dopaminergic neurons, excitatory (glutamate) and inhibitory (GABA) VTA-neurons have also been described: these neurons may also modulate and contribute to shape the final dopaminergic response, which is critical for memory formation. However, given the large amount of information that is handled daily by our brain, it is essential that irrelevant information be deleted. Recently, apart from the well-established role of dopamine (DA) in learning, it has been shown that DA plays a critical role in the intrinsic active forgetting mechanisms that control storage information, contributing to the deletion of a consolidated memory. These new insights may be instrumental to identify therapies for those disorders that involve memory alterations.
... The identification of full length genes, in the present study, under pathways specific to the brain points out the importance of neurotransmission for survival the survival of fish under migration, such as for nervous system (synaptic vesicle cycle, long-term potentiation, dopaminergic synapse, neurotrophin signaling pathway and retrograde endocannabinoid signaling) and development and regeneration (Axon guidance) and under environmental adaptation (Circadian entrainment). Synaptic vesicle cycle recycling helps in the animals adapt to environmental changes in temperature through the function of the neuron [33], while long-term potentiation forms an integral part of neuron adaptation, which forms basis for learning and memory [34]. The role of the fish brain in response to hypo-osmotic stress through the specific signalling events have been reported by several authors [10]. ...
Article
Full-text available
Background Full length transcriptomes, achieved through long-read sequencing, along with the isoform analysis can reveal complexities in the gene expression profiles, as well as annotate the transcriptomes of non-model organisms. Methods and result Full length transcripts of brain transcriptome of Tenualosa ilisha, Hilsa shad, were generated through PacBio single molecule real-time sequencing and were characterized. A total of 8.30 Gb clean reads were generated, with PacBio RSII, which resulted in 57,651 high quality consensus transcripts. After removing redundant reads, a total of 19,220 high-quality non-redundant transcripts and 17,341 full length ORF transcripts were classified to 7522 putative ortholog groups. Genes involved in various neural pathways were identified. In addition, isoform clusters and lncRNAs were discovered, along with Hilsa specific transcripts with coding frames and 29,147 SSRs in 944 transcripts (1141 annotated). Conclusion The present study provided, for the first time, a comprehensive view of the alternative isoforms of genes and transcriptome complexity in Hilsa shad brain and forms a rich resource for functional studies in brain of this anadromous fish.
Article
Chronic pain and alcohol use disorder (AUD) are highly comorbid, and patients with chronic pain are more likely to meet the criteria for AUD. Evidence suggests that both conditions alter similar brain pathways, yet this relationship remains poorly understood. Prior work shows that the anterior insular cortex (AIC) is involved in both chronic pain and AUD. However, circuit-specific changes elicited by the combination of pain and alcohol use remain understudied. The goal of this work was to elucidate the converging effects of binge alcohol consumption and chronic pain on AIC neurons that send projections to the dorsolateral striatum (DLS). Here, we used the Drinking-in-the-Dark (DID) paradigm to model binge-like alcohol drinking in mice that underwent spared nerve injury (SNI), after which whole-cell patch-clamp electrophysiological recordings were performed in acute brain slices to measure intrinsic and synaptic properties of AIC→DLS neurons. In male, but not female, mice, we found that SNI mice with no prior alcohol exposure consumed less alcohol compared with sham mice. Electrophysiological analyses showed that AIC→DLS neurons from SNI-alcohol male mice displayed increased neuronal excitability and increased frequency of miniature excitatory postsynaptic currents. However, mice exposed to alcohol prior to SNI consumed similar amounts of alcohol compared with sham mice following SNI. Together, our data suggest that the interaction of chronic pain and alcohol drinking have a direct effect on both intrinsic excitability and synaptic transmission onto AIC→DLS neurons in mice, which may be critical in understanding how chronic pain alters motivated behaviors associated with alcohol.
Article
Full-text available
Introduction The insular cortex is involved in multiple physiological processes including working memory, pain, emotion, and interoceptive functions. Previous studies have indicated that the anterior insular cortex (aIC) also mediates interoceptive attention in humans. However, the exact cellular and physiological function of the aIC in the regulation of this process is still elusive. Methods In this study, using the 5-choice serial reaction time task (5-CSRTT) testing paradigm, we assessed the role of the aIC in visuospatial attention and impulsiveness in mice. Results The results showed a dramatic activation of c-Fos in the aIC CaMKIIα neurons after the 5-CSRTT procedure. In vivo fiber photometry revealed enhanced calcium signaling in aIC CaMKIIα neurons when the mice responded correctly. In addition, chemogenetic suppression of aIC CaMKIIα neurons led to increased incorrect responses within the appropriate time. Importantly, pharmacological activation of aIC CaMKIIα neurons enhanced their performance in the 5-CSRTT test. Discussion These results provide compelling evidence that aIC CaMKIIα neurons are essential for the modulation of attentional processing in mice.
Article
Full-text available
Gustatory cortical (GC) single-neuron taste responses reflect taste quality and palatability in successive epochs. Ensemble analyses reveal epoch-to-epoch firing rate changes in these responses to be sudden, coherent transitions. Such nonlinear dynamics suggest that GC is part of a recurrent network-that GC produces these dynamics in concert with other structures. basolateral amygdala (BLA), which is reciprocally connected to GC and central to hedonic processing, is a strong candidate partner for GC: BLA taste responses evolve on the same general "clock" as GC; furthermore, inhibition of activity in the BLA➜GC pathway degrades the sharpness of GC transitions. These facts motivate, but do not test, our over-arching hypothesis that BLA and GC act as a single, co-modulated network during taste processing. Here we provide just this test of simultaneous (BLA and GC) extracellular taste responses in female rats, probing the multi-regional dynamics of activity in these regions to directly test whether BLA and GC responses contain coupled dynamics. We show that BLA and GC response magnitudes covary across trials and within single responses, and that changes in BLA-GC LFP phase-coherence are epoch-specific. Such classic coherence analyses, however, obscure the most salient facet of BLA-GC coupling: sudden transitions in and out of the epoch known to be involved in driving gaping behavior happen near-simultaneously in the two regions, despite huge trial-to-trial variability in transition latencies. This novel form of inter-regional coupling, which we show is easily replicated in model networks, suggests collective processing in a distributed neural network.SIGNIFICANCE STATEMENT:There has been little investigation into real-time communication between brain regions during taste processing, a fact reflecting the dominant belief that taste circuitry is largely feedforward. Here, we perform an in-depth analysis of real-time interactions between gustatory cortex (GC) and basolateral amygdala (BLA) in response to passive taste deliveries, using both conventional coherence metrics and a novel methodology which explicitly considers trial-to-trial variability and fast single-trial dynamics in evoked responses. Our results demonstrate that BLA-GC coherence changes as the taste response unfolds, and that BLA & GC specifically couple for the sudden transition into (and out of) the behaviorally-relevant neural response epoch, suggesting (though not proving): 1) recurrent interactions; subserving the dyad's function as 2) a putative attractor network.
Article
Full-text available
To survive in an ever-changing environment, animals must detect and learn salient information. The anterior insular cortex (aIC) and medial prefrontal cortex (mPFC) are heavily implicated in salience and novelty processing, and specifically, the processing of taste sensory information. Here, we examined the role of aIC-mPFC reciprocal connectivity in novel taste neophobia and memory formation, in mice. Using pERK and neuronal intrinsic properties as markers for neuronal activation, and retrograde AAV (rAAV) constructs for connectivity, we demonstrate a correlation between aIC-mPFC activity and novel taste experience. Furthermore, by expressing inhibitory chemogenetic receptors in these projections, we show that aIC-to-mPFC activity is necessary for both taste neophobia and its attenuation. However, activity within mPFC-to-aIC projections is essential only for the neophobic reaction but not for the learning process. These results provide an insight into the cortical circuitry needed to detect, react to- and learn salient stimuli, a process critically involved in psychiatric disorders.
Article
Full-text available
We investigated the role of the ventromedial prefrontal cortex (vmPFC) in extinction of conditioned taste aversion (CTA) by microinfusing a protein synthesis inhibitor or N-methyl-d-asparate (NMDA) receptors antagonist into the vmPFC immediately following a non-reinforced extinction session. We found that the protein synthesis blocker anisomycin, but not the NMDA receptors antagonist D,L-2-amino-5-phosphonovaleric acid, impaired CTA extinction in the vmPFC. Anisomycin microinfusion into vmPFC had no effect on CTA acquisition and by itself did not induce CTA. These findings show the necessary role functional protein synthesis is playing in the vmPFC during the learning of CTA extinction.
Article
Full-text available
Conducted 4 experiments of relevance to a conditioned attention theory of latent inhibition (LI) phenomenon, using 162 Charles River and 56 Sprague-Dawley male albino rats. The conditioning suppositions of the theory predict that the addition of a 2nd stimulus in a conditioning relationship to the preexposed stimulus should maintain attention to that stimulus and thus attenuate LI. Exps I and II demonstrated this effect in the context of lick-suppression conditioning. It is further supposed that S control of the presentation of the preexposed stimulus should maintain attention to that stimulus and thereby reduce LI. This was demonstrated in Exps III and IV, also in lick-suppression conditioning. (22 ref) (PsycINFO Database Record (c) 2012 APA, all rights reserved)
Article
Full-text available
The goal of this book was to call the attention of neuroscientists to the phenomenon of conditioned taste aversion (CTA), the study of which may advance the present neurobiology of learning and memory. Past research has shown that CTA has a number of unusual properties that contrast with the basic tenets of traditional learning theory, and more recent research had advanced the understanding of the neural mechanisms of CTA at the system level and opened up possibilities of their analysis at the molecular and cellular levels. Chapter 1 gives a brief survey of the history and present state of CTA research. Chapter 2 describes the basic methodology and terminology of behavioral tests. Chapters 3–8 review in detail the experimental evidence obtained by irreversible and reversible lesions and by pharmacological, electrophysiological, immunoreactivity, and transplantation studies. Chapter 9 points out the pivotal problems of CTA research and attempts to predict its main direction in the near future. (PsycINFO Database Record (c) 2012 APA, all rights reserved)
Article
Immediate early gene (IEG) expression is dynamically regulated in brain neurons, in response to natural activity, and linked to neural plasticity. Because of the demonstrated importance of synaptic activity in neuronal function and development, it is anticipated that IEGs contribute to adult and developmental cortical plasticity. The IEG response includes proteins that function as transcription factors, as well as enzymes and growth factor proteins that can directly modify neuronal and synaptic function. An understanding of the mechanistic contribution of the IEG response to neuronal plasticity will provide information not only about basic developmental processes, and their disturbances in mental retardation, but may also contribute to new diagnostic and therapeutic approaches for these developmental disorders. (C) 1999 Wiley-Liss. Inc.
Article
Many forms of long-lasting behavioral and synaptic plasticity require the synthesis of new proteins. For example, long-term potentiation (LTIP) that endures for more than an hour requires both transcription and translation. The signal-transduction mechanisms that couple synaptic events to protein translational machinery during long-lasting synaptic plasticity, however, are not well understood. One signaling pathway that is stimulated by growth factors and results in the translation of specific mRNAs includes the rapamycin-sensitive kinase mammalian target of rapamycin (mTOR, also known as FRAP and RAFT-1). Several components of this translational signaling pathway, including mTOR, eukaryotic initiation factor-4E-binding proteins 1 and 2, and eukaryotic initiation factor-4E, are present in the rat hippocampus as shown by Western blot analysis, and these proteins are detected in the cell bodies and dendrites in the hippocampal slices by immunostaining studies. In cultured hippocampal neurons, these proteins are present in dendrites and are often found near the presynaptic protein, synapsin I. At synaptic sites, their distribution completely overlaps with a postsynaptic protein, PSD-95. These observations suggest the postsynaptic localization of these proteins. Disruption of mTOR signaling by rapamycin results in a reduction of late-phase LTP expression induced by high-frequency stimulation; the early phase of LTIP is unaffected. Rapamycin also blocks the synaptic potentiation induced by brain-derived neurotrophic factor in hippocampal slices. These results demonstrate an essential role for rapamycin-sensitive signaling in the expression of two forms of synaptic plasticity that require new protein synthesis. The localization of this translational signaling pathway at postsynaptic sites may provide a mechanism that controls local protein synthesis at potentiated synapses.
Article
The N-methyl-D-aspartate (NMDA) receptor subserves synaptic glutamate-induced transmission and plasticity in central neurons. The yeast two-hybrid system was used to show that the cytoplasmic tails of NMDA receptor subunits interact with a prominent postsynaptic density protein PSD-95. The second PDZ domain in PSD-95 binds to the seven-amino acid, COOH-terminal domain containing the terminal tSXV motif (where S is serine, X is any amino acid, and V is valine) common to NR2 subunits and certain NR1 splice forms. Transcripts encoding PSD-95 are expressed in a pattern similar to that of NMDA receptors, and the NR2B subunit co-localizes with PSD-95 in cultured rat hippocampal neurons. The interaction of these proteins may affect the plasticity of excitatory synapses.
Article
Immediate early gene (IEG) expression is dynamically regulated in brain neurons, in response to natural activity, and linked to neural plasticity. Because of the demonstrated importance of synaptic activity in neuronal function and development, it is anticipated that IEGs contribute to adult and developmental cortical plasticity. The IEG response includes proteins that function as transcription factors, as well as enzymes and growth factor proteins that can directly modify neuronal and synaptic function. An understanding of the mechanistic contribution of the IEG response to neuronal plasticity will provide information not only about basic developmental processes, and their disturbances in mental retardation, but may also contribute to new diagnostic and therapeutic approaches for these developmental disorders. MRDD Research Reviews 1999;5:41–50. © 1999 Wiley-Liss, Inc.
Article
Abstract We studied the role of protein kinase C (PKC) and protein kinase A (PKA) in mediating learning-related long lasting reduction of the post-burst after-hyperpolarization (AHP) in cortical pyramidal neurons. We have shown previously that pyramidal neurons in the rat piriform (olfactory) cortex from trained (TR) rats have reduced post-burst AHP for 3 days after odour-discrimination learning, and that this reduction is due to decreased conductance of calcium-dependent potassium current. In the present study, we examined whether this long-lasting reduction in AHP is mediated by second messenger systems. The broad-spectrum kinase inhibitor, H7, increased the AHP in neurons from TR rats, but not in neurons from pseudo-trained (pseudo-TR) and naive rats. Consequently, the difference in AHP amplitude between neurons from TR and control animals was diminished. This effect was also obtained by application of the specific PKC inhibitor, GF-109203x. The PKC activator, 1-Oleoyl-2-acetyl-sn-glycerol (OAG), significantly reduced the AHP in neurons from naive and pseudo-TR rats, but not in neurons from TR rats, so that the difference between the groups was abolished. The PKA-specific inhibitor, H-89, increased the AHP in neurons from all groups to a similar extent, and the difference in AHP amplitude between neurons from TR rats and neurons from controls was maintained. We suggest that while the post-burst AHP in piriform cortex pyramidal neurons is modulated by both PKC and PKA, a PKC-dependent process maintains the learning-related reduction of the AHP in these cells.