Abstract
A major component of consolidation theory holds that protein synthesis is required to produce the synaptic modification needed for long-term memory storage. Protein synthesis inhibitors have played a pivotal role in the development of this theory. However, these commonly used drugs have unintended effects that have prompted some to reevaluate the role of protein synthesis in memory consolidation. Here we review the role of protein synthesis in memory formation as proposed by consolidation theory calling special attention to the controversy involving the non-specific effects of a group of protein synthesis inhibitors commonly used to study memory formation in vivo. We argue that molecular and genetic approaches that were subsequently applied to the problem of memory formation confirm the results of less selective pharmacological studies. Thus, to a certain extent, the debate over the role of protein synthesis in memory based on interpretational difficulties inherent to the use of protein synthesis inhibitors may be somewhat moot. We conclude by presenting avenues of research we believe will best provide answers to both long-standing and more recent questions facing field of learning and memory.
Keywords: Protein synthesis, RNA synthesis, Consolidation, Memory, Amnesia, Reconsolidation, Protein synthesis inhibitor, Sleep, Gene expression, Transcription, Translation, Knockout, Synaptic, Systems, Side effects, Rescue, Hippocampus, Learning, LTP, cAMP, CREB, NMDA, Tagging, Post-translational modification
1. Introduction
I sometimes feel, in reviewing the evidence on the localization of the memory trace, that the necessary conclusion is that learning is just not possible (Lashley, 1950).
One of the most puzzling questions facing psychologists and neurobiologists alike is one that was posed centuries ago: what is the nature of memory? How does wakeful experience alter neural circuits within the brain in such a precise and meaningful way that even decades later we are able to invoke a remarkably detailed percept of our own history? Indeed, the formation of cognitive associations between external stimuli or between our actions and their consequences can be demonstrated with relative ease. However, it is considerably more difficult to causally connect cellular and molecular events to the instantiation of such associations. Nevertheless, we are now equipped with sophisticated molecular and genetic techniques that afford us the opportunity to probe deeper than ever before into the molecular underpinnings of memory.
In the first section of this review, we examine the emergence of consolidation theory—the idea that memories are stabilized over time—recalling several important findings that were seminal to its development and continuing evolution. We then examine the basis of the long-standing debate regarding the validity of a major tenet of consolidation theory: that new proteins must be synthesized to stabilize newly acquired memories. Indeed, this debate, although ignored by many, has never been resolved to the satisfaction of some. We then briefly summarize the remarkable progress that has been made in understanding consolidation in spite of this debate and present promising new approaches being developed to address some old questions as well as questions that have arisen along the way. Lastly, we conclude by addressing a few alternatives or addendums to consolidation theory that merit consideration.
2. The emergence of consolidation theory: The role of studies using protein synthesis inhibitors
Early empirical forays investigating memory function began in 1878 when Hermann Ebbinghaus introduced the concept of “retroactive interference”. Using sequentially memorized lists of nonsense syllables, Ebbinghaus showed that “forgetting” could be attributed to the interfering effects of subsequently learned matter (Ebbinghaus, 1885), thereby establishing the existence of temporal constraints on memory formation and storage. In 1900, using improved methodology and controls, Müller and Pilzecker confirmed that memory for verbal material was susceptible to disruption if new material was introduced too soon after the initial acquisition period. Thus, they proposed that new memories required a period of “consolidation” to fixate or become resistant to disruption (Lechner, Squire, & Byrne, 1999; Müller & Pilzecker, 1900).
Support for Müller and Pilzecker’s consolidation theory came a half-century later when it was observed that memory in rats could be retroactively disrupted by applying an electroconvulsive shock near the time of training (Duncan, 1949) or through head injuries involving the hippocampus and related structures (Russell & Nathan, 1946). In both forms of retrograde amnesia, memory loss varies inversely with the age of the memory where new memories are more susceptible to disruption. The amnesia described by Russell and Nathan, which can extend for years prior to the actual neural insult, led to the hypothesis that memories are formed and stored in the hippocampus temporarily but are then transferred to distal cortical sites for permanent storage. This relatively slow process is now referred to as systems consolidation (McGaugh, 2000; Squire & Bayley, 2007). This is in contrast to the relatively faster processes of stabilization revealed by verbal interference, electroconvulsive shock, and pharmacological experiments (described below) thought to occur on a cellular or synaptic or level (McGaugh, 1966). Parenthetically, the precise time course during which synaptic consolidation occurs is unclear but has been reported to range anywhere from 500 ms to hours depending on the type of memory being examined, the training procedures, and the amnestic agent used to probe memory (Miller & Matzel, 2006).
This review focuses primarily on the role of protein synthesis during synaptic consolidation largely because the vast majority of cellular and molecular research has targeted the more accessible processes occurring immediately after novel learning situations. However, the existence of systems consolidation must also dictate to an equal extent how we envision and study memory storage over extended periods of time (Frankland & Bontempi, 2005).
Another major step in the evolution of consolidation theory that changed the way the field would conceptualize the consolidation of memory on a cellular level also occurred in 1949 when Hebb introduced his “dual-trace hypothesis” of memory formation. Hebb proposed that reverberation of activity within assemblies of neurons was the essence or trace of short-term memory and that if maintained long enough some growth processes at the level of the synapse could lead to long-term memory (Hebb, 1949). Indeed, disruptions in neuronal reverberation was seen as an attractive explanation of the mechanism by which retrograde amnesia might occur (Glickman, 1961; McGaugh, 1999; Misanin, Miller, & Lewis, 1968; Schneider & Sherman, 1968).
Further support for consolidation theory was offered in the late 1950s, after Scoville and Milner described the memory deficits experienced by the famous patient H.M. After bilateral resection of the medial structures of the temporal lobe to treat epilepsy, it was evident that H.M. had severe short-term memory deficits, unable to form new hippocampus-dependent long-term memories (Scoville & Milner, 1957). Importantly, when tested in delayed matching and delayed comparison tasks, H.M.’s performance worsened as the delays increased leading Milner to conclude, in support of Hebb’s view of consolidation, that a distinction exists between the initial processing of memory that decays rapidly and a later, secondary process that is responsible for long-term storage of information (Milner, 1972).
Yet, while great progress was being made in understanding memory from observations from both human and animal studies, direct evidence of the mechanisms by which experience produced long-lasting neural changes was still lacking. This would all change when, in considering how memory traces might be instantiated within neural circuitry on a molecular level Monné, as well as Katz and Halstead, hypothesized that protein molecules were somehow required (Katz & Kalstead, 1950; Monné, 1948; Sutton & Schuman, 2006). At last, after a decade or so of speculation, Flexner and colleagues demonstrated memory for a discriminative avoidance task (performed in a Y-maze) could be disrupted in mice using the protein synthesis inhibitor puromycin (Flexner, Flexner, Stellar, De La Haba, & Roberts, 1962; Flexner, Flexner, & Stellar, 1963). Additionally, some cortical specificity in memory formation was evident in that only bilateral temporal injections affecting the hippocampus and adjacent temporal cortex consistently disrupted memory for recently acquired memories. Consistent with the notion of systems consolidation, temporal infusions in combination with infusions more rostral and caudal were required to disrupt more remote memories (e.g.,11–43 days old), indicating the establishment of a more distributed trace over time (Flexner et al., 1963). A flurry of reports investigating the role of a variety of protein synthesis inhibitors on different memory tasks soon followed (Agranoff, 1967; Agranoff, Davis, & Brink, 1965; Agranoff & Klinger, 1964; Barondes & Cohen, 1966; Barondes & Cohen, 1967; Davis & Squire, 1984; Dudai, 2004; Flexner & Flexner, 1966; Flexner, Flexner, & Roberts, 1966; Flexner, Flexner, & Stellar, 1965; McGaugh, 2000). It should also be noted that other pharmacological treatments were found to facilitate rather than disrupt memory (Doty & Doty, 1966; Hunt & Krivanek, 1966; McGaugh & Petrinovich, 1965; Pare, 1961; Stratton & Petrinovich, 1963).
In general, it was found that protein synthesis inhibitors, like electroconvulsive shock, were most efficacious when administered around the time of training (but see below for exceptions) having little effect on established memories. Therefore, consolidation has historically been assumed to occur but once, stabilizing the engram permanently. However, this notion too is currently under scrutiny for processes such as multiple “waves” of consolidation (Bourtchouladze et al., 1998; Freeman, Rose, & Scholey, 1995; Scholey, Rose, Zamani, Bock, & Schachner, 1993; Stork & Welzl, 1999) and reconsolidation (Alberini, 2005; Dudai, 2006; Dudai & Eisenberg, 2004; Nader, 2003; Tronson & Taylor, 2007).
Given the central dogma of molecular biology at that time, specifically that DNA is transcribed to synthesize RNA which is translated to synthesize protein, the ability of RNA synthesis inhibitors to disrupt memory quickly followed the Flexners’ initial studies using protein synthesis inhibitors. Early studies using the RNA synthesis inhibitor actinomycin D produced mixed results. For example, actinomycin D successfully impaired memory in goldfish when infused immediately after training but not after a 1-h delay (Agranoff, Davis, Casola, & Lim, 1967) suggesting the synthesis of new proteins subsequent to learning-induced transcription may be important in memory formation. However, in rodents, systemic administration of actinomycin D was extremely toxic complicating the interpretation of these studies (Squire & Barondes, 1970). Nevertheless, later studies would provide a better demonstration of the requirement of transcription in memory formation as summarized below (Igaz, Vianna, Medina, & Izquierdo, 2002).
Most importantly these early studies demonstrating the involvement of de novo protein synthesis in memory formation made tangible the notion that the the physical basis of memory lies in the learning-related growth or remodeling of synaptic connections in a protein synthesis-dependent fashion. However, this notion was not universally accepted.
3. Putting the breaks on translation: Non-specific effects of protein synthesis inhibitors on memory
By providing the first experimental evidence of a protein synthesis-dependent stage of memory formation, the Flexners catalyzed the incorporation of protein synthesis into the very definition of consolidation. Ironically, it was also the Flexners who incited the ongoing debate regarding the actual role of de novo protein synthesis in memory formation. Adding to the irony is that this theoretical “about face” was sparked after Flexner and colleagues demonstrated that an injection of saline could recover memory thought to be destroyed by puromycin (Flexner & Flexner, 1967). Although the mechanism by which saline restored memory was a mystery, it was nevertheless concluded that puromycin only blocked the expression of memory by acting on cellular mechanisms not related to protein synthesis. However, it was soon discovered that saline was only able to restore memory when puromycin was administered 24 h or more after training; puromycin infusions given immediately before or after training resulted in memory impairments that could not be restored (Flexner & Flexner, 1968a).
3.1. Effects of altered catecholamine function and pharmacological rescue of amnesia
Suspicions that protein synthesis inhibitors had serious side effects grew after it was shown that puromycin inhibited the increase in norepinephrine synthesis normally observed after stimulation of the hypogastric nerve of the vas deferens in the guinea-pig (Weiner & Rabadjija, 1968). Reduced norepinephrine levels were also known to temporarily impair performance on conditioned avoidance responses (Schoenfeld & Seiden, 1969; Seiden & Peterson, 1968). An experiment performed by Flexner and colleagues then tested a variety of drugs on their ability to restore memory in mice that were injected bitemporally with puromycin 24 h after training in the Y-maze. Memory in a few instances was found to be at least partially restored by drugs that act on the adrenergic system (e.g., reserpine or L-DOPA). Therefore, the authors hypothesized that the adsorption of puromycin to adrenergic sites might underlie memory impairments rather than the inhibition of translation (Roberts, Flexner, & Flexner, 1970). It should be noted that puromycin in this study was not administered during the critical period immediately after acquisition during which memory is thought to depend most on protein synthesis. Thus, these drug treatments could have been acting to enhance some aspect of less severely disrupted memories. Furthermore, it should be noted that the drug treatments used in the above study were administered to a very limited number of mice and killed as many as had experienced full or partial memory restorations.
Subsequent evidence demonstrated that a variety of protein synthesis inhibitors decreased tyrosine hydroxylase activity, thereby altering endogenous catecholamine levels (Flexner & Goodman, 1975; Flexner, Serota, & Goodman, 1973). Of interest, is a study by Flexner and colleagues that examined the effect of infusions of the predominately α1-adrenergic agonist metaraminol on acetoxycycloheximide-induced amnesia 24 h post-training in rats. Bitemporal injections of acetoxycycloheximide were administered 5 h prior to training in a Y-maze, a procedure known to only cause transient amnesia. Metaraminol boosted performance levels when administered 0.5 h before training, 0–2 h after training, or 0.5–2 h before testing. Metaraminol infusions administered from 2.5–21.5 h after training had no significant effect. The authors suggest that acetoxycycloheximide produced amnesia not by inhibiting protein synthesis but by lowering the amount of norepinephrine available at the time of training (Serota, Roberts, & Flexner, 1972). Indeed, this study has been cited (Routtenberg & Rekart, 2005) as a clear demonstration that protein synthesis inhibitors act to impair memory through mechanisms other than by inhibition of translation. However, the performance improvements observed after metaraminol infusions around the time of training or testing imply that encoding and retrieval mechanisms were enhanced, respectively. The lack of effect of metaraminol in the 2.5–21.5 h group is consistent with the fact that the potency of the drug had diminished enough over the 2.5 h delay between the end of drug administration and memory testing such that retrieval was no longer enhanced. Indeed, it could be argued that this study actually supports consolidation theory inasmuch as the drug had no effects after administration between 2.5 and 21.5 h—a period of time when receptor agonism might not be expected to affect memory consolidation. Importantly, the inhibition of protein synthesis is known not to impair encoding or retrieval (Davis & Squire, 1984). Thus, it seems that metaraminol did not rescue amnesia by reversing the effects of acetoxycycloheximide but overcame them perhaps by enhancing the encoding and retrieval of the weakly consolidated memory. Conceptually, these studies suggest that a single agent might affect disparate cellular processes (e.g., protein synthesis or neurotransmitter function) by interfering with temporally and biochemically distinct aspects of memory processing (e.g., encoding versus consolidation or retrieval). In the case of puromycin, consolidation seems to be impaired by protein synthesis inhibition and retrieval through interference with neurotransmitter function. However, this lesson is often overlooked even today.
Others involved in this debate argued for the role of translation in consolidation, perhaps more convincingly. For example, Squire and colleagues found that repeated injections of α-methyl-p-tyrosine, an inhibitor of tyrosine hydroxylase, left long-term memory unaffected in spite of the fact that brain tyrosine hydroxylase activity was inhibited to levels equal to or surpassing that caused by doses of cycloheximide or anisomycin used to impair memory (Squire, Kuczenski, & Barondes, 1974). A further study by Flood and colleagues compared the effects of anisomycin with the effects of three catecholamine synthesis inhibitors across seven different behavioral paradigms. They reported that pre-trial infusions of any of the four drugs impaired passive avoidance learning but only when training was weak. However, only anisomycin impaired learning when infused post-trial during the consolidation period (Flood et al., 1986).
Also relevant to this discussion is a study conducted by Squire and Barondes where they examined the effects of cycloheximide on memory for a shock-escape object discrimination task at a time when 95% of brain protein synthesis was inhibited (1 min, ~95% inhibition at time of training) or after inhibition had been established for significant period of time (118 min, ~70% inhibition at time of training). Mice injected 1 min before training acquired the task but did not retain the memory when tested the next day. In contrast, mice injected with the same dose of cycloheximide 118 min before training (or 10 min after training) exhibited normal retention (Squire & Barondes, 1976). Since cycloheximide elevates tryrosine levels by 50% from 30 min to 4 h after injection (Flexner et al., 1973) then the accumulation would be expected to be greater in the 118 min group and therefore disrupt memory to a greater degree relative to the 1 min group. However, only the 1 min treatment caused amnesia for the task. Similarly, since alterations in tyrosine hydroxylase activity and catecholamine synthesis develops over 2–3 h after injection the same argument can be made that these effects should be greatest in the group that produced no memory disruptions. The same study also suggests that protein synthesis inhibitors do not cause amnesia by blocking the replacement of constitutive proteins (even those with a half-life of as little as 11 min) but do so by blocking the induction of new proteins needed for memory formation. Taken together, these findings demonstrate that altered tyrosine hydroxylase activity and catecholamine levels, at least on their own, cannot explain the amnesia induced by protein synthesis inhibitors.
Since the Flexners’ early memory restoration studies, a variety of drugs (e.g., amphetamine, nicotine, fluoxetine, etc.) have been found to enhance performance in memory tasks even while animals are under the influence of various protein synthesis inhibitors. Because they do so without diminishing levels of protein synthesis inhibition, the apparent “rescue” of amnesia by these pharmacological agents has been cited as further evidence against the requirement of de novo protein synthesis in memory formation. However, drug treatments that seem to rescue the memory impairing effects of protein synthesis inhibitors could do so by acting on parallel cellular mechanisms that promote protein synthesis thereby overcoming the effects of protein synthesis inhibitors (Davis & Squire, 1984; Gold, 2006) or by modulating other mechanisms on which memory formation depends (e.g., metaraminol). Some drugs enhance memory by mimicking the brain’s natural response to arousing or stressful situations (i.e., through the release of epinephrine, glucose, or corticosterone) (Gold, 2003; Gold, 2005; McGaugh, 1983). Importantly, the fact that certain lesions (e.g., adrenalectomy) only prevent the endogenous enhancement of memory (i.e., leaving basic memory intact) demonstrates the notion that certain drugs do not truly rescue amnesia but act to enhance the residual memory that typically exists after various amnestic treatments (Gold, 2006; Roozendaal, Carmi, & McGaugh, 1996; Roozendaal & McGaugh, 1996a; Roozendaal & McGaugh, 1996b). Alternatively, some drugs (in particular drugs of abuse) might act to enhance performance by over-stimulating the very same signaling pathways recruited during experience-dependent learning leading to a type of maladaptive “hijacking” of learning and memory systems (Flood & Cherkin, 1987; Goodman, Flexner, & Flexner, 1975; Hyman, Malenka, & Nestler, 2006; Kalivas, 2005; Kelley, 2004).
3.2. Mechanisms of action and related toxicity of protein synthesis inhibitors
Adding to doubts over the specificity of protein synthesis inhibitors on translation is the fact that they are all significantly toxic at levels needed to cause amnesia. Yet, the antibiotics puromycin, anisomycin, ementine, cycloheximide and acetoxycycloheximide continue to be widely used to inhibit translation in vivo (Davis & Squire, 1984; Stork & Welzl, 1999; Uphouse, MacInnes, & Schlesinger, 1974). The following section summarizes some of the issues surrounding mechanisms of translation inhibition and toxicity. Puromycin acts by incorporating into the growing peptide chain causing premature release from the ribosomal complex and the production of abnormal peptidyl–puromycin fragments (Flexner & Flexner, 1968b; Nathans, 1964). Side effects include hippocampal seizures, swelling of mitochondria, and disaggregation of ribosomes (Flood, Rosenzweig, Bennett, & Orme, 1973). Emetine complexes with 80S ribosomal subunits impairing amino acid-tRNA binding. Cycloheximide and acetoxycycloheximide AXM inhibit chain initiation as well as chain elongation by interacting with 60S ribosomal subunits but also impair DNA and RNA synthesis (Gale, Cundliffe, Reynolds, Richmond, & Waring, 1981). Anisomycin, presumably having fewer side effects than other inhibitors (Flood, Bennett, Orme, & Rosenzweig, 1975; Flood et al., 1973; Squire & Barondes, 1974), interferes with protein synthesis by inhibiting chain elongation (Pestka, 1971; Vasquez, 1979). All protein synthesis inhibitors cause visible symptoms of distress (e.g., piloerection, lethargy, diarrhea, and gustatory aversions) when administered systemically in rodents (Davis & Squire, 1984; Squire, Emanuel, Davis, & Deutsch, 1975). Historically, the toxicity of protein synthesis inhibitors has been deemed acceptable as long as the animal’s physical ability to learn or perform is not compromised. Post-training administration of protein synthesis inhibitors followed by behavioral testing the after the effects of the drug have dissipated is a method commonly used to avoid the confounding effects of drug toxicity on memory readouts (but see Hernandez & Kelley, 2004). It should also be noted that the RNA synthesis inhibitor actinomycin D acts by binding DNA at the transcription initiation complex preventing elongation by RNA polymerase (Sobell, 1985). However, use of this drug has been plagued by the fact that it has been shown to cause necrosis and electrical abnormalities in the brain as well as behavioral signs of toxicity that result in death when delivered systemically (Barondes & Jarvik, 1964; Nakajima, 1969; Wetzel, Ott, & Matthies, 1976). The less toxic mRNA synthesis inhibitor 5,6-dichloro-1-β-D-ribofuranosylbenzimidazole (DRB) acts by inhibiting RNA polymerase II and has been successfully used as an alternative to actinomycin D in more recent studies to impair memory (Apergis-Schoute, Debiec, Doyere, LeDoux, & Schafe, 2005; Huang, Martin, & Kandel, 2000; Nguyen, Abel, & Kandel, 1994).
3.3. MAPK activation and apoptosis: Protein synthesis inhibitor induced death or survival?
There is some concern that protein synthesis inhibitors might impair memory via apoptotic cell death. Indeed, protein synthesis inhibitors, especially anisomycin, have long been used as apoptotic agents in cell culture (Shifrin & Anderson, 1999). Anisomycin is thought to cause apoptosis through ribotoxicity or through the activation of MAP kinases (the role of MAPK in learning and memory is discussed in more detail below) (Iordanov et al., 1997). MAPK transduces extracellular signals from tyrosinekinase receptors and guanine nucleotide-binding regulatory proteins (G-protein) coupled receptors to cytoplasmic and nuclear effectors (Chang & Karin, 2001; Schaeffer & Weber, 1999). Three related MAPKs have been identified in neurons: the extracellular signal regulated kinase (ERK), the c-Jun terminal kinase (JNK), and p38 (Obata et al., 2004). ERK signaling cascades mediate cell development, survival and plasticity (Seger & Krebs, 1995; Sweatt, 2001), whereas p38 and JNK, both of which are activated by anisomycin (Eriksson, Taskinen, & Leppa, 2007), respond to cellular stress promoting cell death (Coffey, Hongisto, Dickens, Davis, & Courtney, 2000; Ip & Davis, 1998; Kyriakis & Avruch, 1996; Kyriakis, Woodgett, & Avruch, 1995). However, the exact role of these kinases in cellular survival and death is complex and varies according to cell type, culture conditions, and state of differentiation. For example, JNK activation in differentiated cells causes apoptosis, whereas in undifferentiated cells it leads to differentiation with no adverse effect on cell viability (Heasley et al., 1996; Le-Niculescu et al., 1999; Leppa, Eriksson, Saffrich, Ansorge, & Bohmann, 2001; Xia, Dickens, Raingeaud, Davis, & Greenberg, 1995). With regard to p38, studies suggest that inhibition of p38 signaling can be both lethal and protective (Harada & Sugimoto, 1999; Hong, Qian, Zhao, Bazzy-Asaad, & Xia, 2007; Horstmann, Kahle, & Borasio, 1998; Kharlamov et al., 1995; Maroney et al., 1998; Schaeffer & Weber, 1999; Zheng & Zuo, 2004). Xia and colleagues reported that in PC12 cells that a balance exists between ERK and JNK-p38 in determining whether the cell survives or undergoes apoptosis (Xia et al., 1995). Thus, it seems unclear, at best, how plasticity-related increases in ERK activity in vivo might interact with anisomycin, for example, with regard to the balance of ERK/JNK-p38 signaling and apoptosis. However, two pieces of in vivo evidence exist suggesting that apoptosis is not a major contributor to amnesia induced by protein synthesis inhibitors. First, Lopez-Mascaraque and Price demonstrated in rat cerebral cortex that anisomycin or emetine actually prevented apoptosis in cortical cells at doses used to cause experimental amnesia (Lopez-Mascaraque & Price, 1997). Second, cycloheximide, anisomycin, and emetine have been shown to be effective neuroprotective agents when coadministered with drugs of abuse such as methamphetamine in rats (Finnegan & Karler, 1992) or against hydrogen peroxide induced apoptosis (Gardner et al., 1997). Thus, the effects of protein synthesis inhibitors in vitro may have little to do with their actual effects on the brain.
3.4. Failure to impair long-term memory by protein synthesis inhibition
The resistance of some forms of memory to the effects of protein synthesis inhibitors (Dunn, Gray, & Iuvone, 1977; Flexner & Flexner, 1966; Flexner et al., 1966; Rainbow, Hoffman, & Flexner, 1980) has contributed to suspicions that consolidation does not depend on protein synthesis. Upon closer inspection, however, many of these studies were actually designed to better define the boundary conditions (e.g., training parameters, dosing schedules) in which protein synthesis inhibitors produce amnesia (Cherkin, 1969). For example, Rainbow and colleagues, reported that cycloheximide had no effect when administered to mice 2 h prior to training. However, at 30 min prior to training, when protein synthesis inhibition was significantly greater at the time of training, memory for the same tasks was significantly impaired. Additionally, Flexner’s group failed to produce amnesia consistently in mice after the administration of acetoxycycloheximide 20–24 h post-training, long after the bulk of synaptic consolidation is thought to occur. Yet, this study is still cited as evidence that protein synthesis is not required for consolidation. Finally, in the study by Dunn and colleagues (Dunn et al., 1977), emetine and pactamycin failed to impair memory for a passive avoidance task in mice. A closer inspection of their results reveals that the doses of emetine and pactamycin used resulted in ~50% inhibition of cerebral protein synthesis—far less than typically needed to produce amnesia. Anisomycin and cycloheximide were also tested in this study and were found to produce marked disruptions in memory for the task.
It has been proposed that extinction in mice (Lattal & Abel, 2001), olfactory memories in rats and honeybees (Staubli, Faraday, & Lynch, 1985; Wittstock, Kaatz, & Menzel, 1993), color discrimination in honeybees (Wittstock & Menzel, 1994), and some forms of long-term potentiation (a leading cellular model of memory) (Kurotani, Higashi, Inokawa, & Toyama, 1996) are not affected by protein synthesis inhibitors. However, extinction in other preparations is sensitive to protein synthesis inhibitors (Berman & Dudai, 2001; Eisenberg, Kobilo, Berman, & Dudai, 2003; Quirk, 2004). Interestingly, spontaneous recovery of an extinguished appetitive olfactory memory in honeybees was blocked using emetine (Stollhoff, Menzel, & Eisenhardt, 2005) demonstrating that some form of olfactory memory in the honeybee appears to be sensitive to protein synthesis inhibitors. As Lattal and Abel (2001) point out, differences in the nature of the task, procedural details (e.g., strength of the conditioned stimulus), and molecular differences by which extinction memories might be acquired could account for these contradictory findings.
3.5. Much ado about nothing?
At first glance, the issue seems simple enough: inhibitors of protein synthesis have a range of unintended side effects that could account for experimentally induced amnesia. Why, then, has this debate never been resolved? As we attempted to convey above, the field has been permeated with conflicting results, conclusions based on negative or misinterpreted results, and a reliance on in vitro studies using various undifferentiated, non-neuronal cell lines. Thus, it is important to realize that a simple demonstration of the existence of side effects of translation inhibitors without an understanding of the totality of those effects does not negate the preponderance of evidence in favor of a role for protein synthesis in memory consolidation. However, could it be that critics of the role of de novo protein synthesis in memory formation are right but for the wrong reasons?
4. Synaptic plasticity, long-term memory and protein synthesis: Progress amid the debate
Before we entertain this question, it is important to describe some of the advances the field has achieved since the 1960s and 1970s. The early pharmacological experiments with inhibitors of protein synthesis captured the interest of biochemists and molecular and cellular biologists, providing the inspiration to tackle complex behavioral phenomena including learning and memory. Indeed, a wide range of significant insights into memory storage have since been gained (Kandel, 2001). These cellular and molecular studies have defined the role of transcription and translation in synaptic plasticity and memory storage.
4.1. The discovery of LTP
Some 20 years after Hebb had postulated such a phenomenon, Bliss and Lømo found that a high frequency train of action potentials resulting from stimulation of the perforant path in the rabbit hippocampus led to a long-term potentiation (LTP) of synaptic transmission in the dentate gyrus (Bliss & Lømo, 1973). That activity-dependent synaptic plasticity could be induced for long periods of time by relatively physiological stimuli in a structure known to be involved in memory was extremely significant (Acsady & Kali, 2007; Eccles, 1979; Kullmann & Lamsa, 2007; O’Keefe & Nadel, 1978; Squire, Knowlton, &Musen, 1993). The significance was further realized upon development of the in vitro hippocampal slice preparation (Alger & Teyler, 1976; Lynch, Dunwiddie, & Gribkoff, 1977; Schwartzkroin & Wester, 1975). The slice preparation would revolutionize how years of theoretical supposition regarding the mechanisms of associative memory formation would be explored. For example, the slice preparation made it possible to directly measure and image activity-dependent changes in calcium levels, record from specific neurons, receptors, and channels, and assess the effects of various pharmacological agents on a multitude of cellular processes (e.g., intracellular signaling, transcription, and translation). The key question, of course, is how these data regarding LTP in hippocampal slices relate to behavior. Importantly, a number of correlations and interactions between behavior and LTP have since been demonstrated (Barnes, 1979; Fedulov et al., 2007; Martin, Grimwood, & Morris, 2000; Sharp, McNaughton, & Barnes, 1985; Weisz, Clark, & Thompson, 1984; Whitlock, Heynen, Shuler, & Bear, 2006; Wilson, Willner, Kurz, & Nadel, 1986). Considering the many similarities between LTP and memory, it is now widely held that LTP holds the key to understanding how memories are formed on a cellular and molecular basis. Yet, it should be noted that the slice preparation obviously suffers from the lack of any sort of influence from systems processes.
4.2. Role of protein synthesis in different phases of memory formation and synaptic plasticity
With both in vitro and in vivo models available to study the cellular mechanisms underlying synaptic plasticity and memory, many details of consolidation were rapidly elucidated. As alluded to in our discussion of Hebb and H.M., consolidation theory has come to embrace the idea that memories are formed during at least two phases (Bourtchouladze et al., 1998; DeZazzo & Tully, 1995; Dubnau & Tully, 1998; Ghirardi, Montarolo, & Kandel, 1995; Grecksch & Matthies, 1980; Matthies, 1974; Matthies, 1989). Short-term memory is produced in the first moments after acquisition and lasts minutes to hours, whereas long-term memory is formed during a second phase, lasting from hours to days or longer depending on the organism and type of memory. Correlates in LTP termed early LTP (E-LTP) and late (L-LTP) have also been identified. Short-term memory and E-LTP depend on the post-translational modification of post-synaptic proteins, whereas long-term memory and L-LTP depend on intracellular signaling and the regulation of transcription and translation (Abel & Lattal, 2001; Frey, Krug, Reymann, & Matthies, 1988; Goelet, Castellucci, Schacher, & Kandel, 1986; Huang, 1998; Krug, Lossner, & Ott, 1984; Malenka & Nicoll, 1999; Matthies et al., 1990; Nguyen & Woo, 2003; Reymann & Frey, 2007). An intermediate form of synaptic plasticity or facilitation has also been described in chicks and rats (Allweis, 1991; Frieder & Allweis, 1978; Frieder & Allweis, 1982; Gibbs & Ng, 1979; Rosenzweig, Bennett, Colombo, Lee, & Serrano, 1993), as well as in Aplysia californica (Ghirardi et al., 1995; Stough, Shobe, & Carew, 2006; Sutton & Carew, 2000; Sutton, Masters, Bagnall, & Carew, 2001).
Interestingly, long-term memory in Drosophila melanogaster can be pharmacologically and genetically divided into two types that develop in parallel (Tully, Preat, Boynton, & Del Vecchio, 1994). One form of long-term memory, anesthesia-resistant memory (ARM), is produced after massed or spaced training procedures, and is impaired in radish mutants (Folkers, Waddell, & Quinn, 2006). The other form of long-lasting memory is protein synthesis-dependent, occurs only after spaced training, requires the Drosophila homolog to cyclic adenosine monophosphate (cAMP)-response element-binding protein (CREB), dCREB2, and can be blocked by the protein synthesis inhibitor cycloheximide (Folkers, Drain, & Quinn, 1993; Tully et al., 1994; Yin et al., 1994). Thus, it might be possible that genetically and functionally independent forms of long-term memory also exist in other organisms but have yet to be identified.
The persistence of protein synthesis-dependent changes involved in long-term memory formation may require recurrent rounds or waves of translation of certain proteins (Bekinschtein et al., 2007). Indeed, multiple waves of transcription and translation have been observed during the establishment of long-term facilitation in Aplysia (Barzilai, Kennedy, Sweatt, & Kandel, 1989) and in LTP induction in the hippocampus (Abraham et al., 1993). Furthermore, there are two or more periods during which protein synthesis inhibitors exert their amnesic effects in vivo during the formation of various associative memories (Bourtchouladze et al., 1998; Chew, Vicario, & Nottebohm, 1996; Freeman et al., 1995; Grecksch & Matthies, 1980).
L-LTP (Apergis-Schoute et al., 2005; Frey, Frey, Schollmeier, & Krug, 1996; Nguyen et al., 1994) and long-term memory in a variety of species (Agranoff et al., 1967; Bailey, Kim, Sun, Thompson, & Helmstetter, 1999; Castellucci et al., 1986; Codish, 1971; Crow, Siddiqi, & Dash, 1997; Huang et al., 2000; Schafe et al., 2000) can also be impaired by the inhibition of RNA synthesis. In light of these data, consolidation theory has come to hold that access to the genetic program in the form of activity dependent transcription and the subsequent synthesis of new proteins in the soma is required for memory formation. However, as we learn more of the function of the dendrite in memory formation (Steward & Worley, 2002) the actual “instructive” role of transcription in the nucleus and protein synthesis in the soma needs to be put into the context of the role of dendritic protein synthesis.
4.3. From the outside in: Pathways to protein synthesis
The use of protein synthesis inhibitors paved the way for molecular and genetic dissection of complex behavioral phenomenon such as learning and memory formation. Indeed, the theoretical implications of early pharmacological studies using these inhibitors, spurred the molecular and genetic investigation into learning impairments observed in the Drosophila dunce mutant (Dudai, Jan, Byers, Quinn, & Benzer, 1976) leading to the realization of the importance of cyclic adenosine monophosphate/protein kinase A (cAMP/PKA)-mediated signaling in learning and memory (Byers, Davis, & Kiger, 1981; Kalderon & Rubin, 1988). Similarly, the connection between protein synthesis and memory formation was applied to Aplysia where it was found that protein synthesis-dependent synaptic facilitation was required for sensitization of the gill and siphon withdrawal reflex (Montarolo et al., 1986). Further study revealed a critical role for the transcription factor cAMP response element binding protein (CREB) in this form of plasticity (Dash, Hochner, & Kandel, 1990) and rapamycin sensitive protein synthesis (Casadio et al., 1999) where LTP studies largely elucidated the critical role of N-methyl-D-aspartate (NMDA) receptors in subsequent signaling and the stimulation of immediate early gene expression (Morgan & Curran, 1988; Worley, Cole, Saffen, & Baraban, 1990) and new protein synthesis resulting in synaptic plasticity and long-term memory (Bliss & Collingridge, 1993; Kim, DeCola, Landeira-Fernandez, & Fanselow, 1991; Morris, 1989; Nakazawa, McHugh, Wilson, & Tonegawa, 2004; Rao & Finkbeiner, 2007; Verkhratsky & Kirchhoff, 2007).
These seminal studies initiated much research focused on understanding how learning-related environmental information is transduced to affect synaptic change. It is now known that, once activated by coincident pre-synaptic release of glutamate and post-synaptic depolarization, NMDA receptors permit calcium entry into the cell triggering a range of intracellular signaling cascades some of which result in plasticity-related gene transcription and translation (Collingridge, Kehl, & McLennan, 1983; Cotman et al., 1989; Kasten, Fan, & Schulz, 2007). NMDA receptor signaling through the cAMP/PKA/CREB pathway has been extensively studied (Blokland, Schreiber, & Prickaerts, 2006; Dash, Moore, Kobori, & Runyan, 2007; Skoulakis & Grammenoudi, 2006; Wu, Zhou, & Xiong, 2007). Briefly, the influx of calcium stimulates calcium binding protein Ca2+/calmodulin to increase the production of cAMP by adenylyl cyclases (Eliot, Dudai, Kandel, & Abrams, 1989). cAMP can also be synthesized by adenylyl cyclases linked to G-proteins upon the binding of non-glutamate transmitters (e.g., dopamine) and hormones to their receptors (Tang & Gilman, 1991). cAMP then activates PKA which in turn activates CREB (Impey et al., 1998; Roberson et al., 1999; Vanhoutte et al., 1999a). CREB can also be activated though calcium signaling through the mitogen-activated protein kinase/extracellular signal-regulated kinase (MAPK/ERK) pathway where a number of important kinases including CaMKII, phosphatidylinositol 3-kinase (PI3-kinase), and protein kinase C (PKC) are involved in signaling from the NMDA receptor to MAPK/ERK (Davis & Laroche, 2006; Miyamoto, 2006; Soderling & Derkach, 2000; Wang & Storm, 2003). Thus, MAPK/ERK is thought to coordinate cross-talk between several different kinase signaling pathways, providing an addition level of control over CREB-mediated transcription during memory formation (Impey et al., 1998; Sweatt, 2001; Vanhoutte et al., 1999b; Wang, Fibuch, & Mao, 2007).
CREB along with is its co-activator CREB binding protein (CBP) is then primed to regulate the expression of a variety of plasticity-related genes with CRE response elements in their promoter region (Bacskai et al., 1993; Hagiwara, Shimomura, Yoshida, & Imaki, 1996). A number of these genes are immediate-early genes and act as transcription factors regulating the expression of “effector” or late response genes. Ultimately, it is the protein products of effector genes that compose the structural components needed for the growth and/or stabilization of synapses needed in memory formation (Bailey, Bartsch, & Kandel, 1996; Milner, Squire, & Kandel, 1998; Steward, Wallace, Lyford, & Worley, 1998; Yin & Tully, 1996). Recent work also suggests that CREB/CBP is able to direct long-term epigenetic changes to structure of chromatin by acetylating histones and that this pattern of histone acetylation could mediate memory in this fashion by providing the transcriptional machinery access to plasticity-related genes (Korzus, Rosenfeld, & Mayford, 2004; Levenson & Sweatt, 2006; Vecsey et al., 2007). Indeed, mice with mutations in various plasticity-related genes, in particular those with altered PKA, CREB, or CBP activity, demonstrate marked disruptions in learning and memory (Abel et al., 1997; Bourtchouladze et al., 1994; Bourtchouladze et al., 2003; Brandon et al., 1995; Korzus et al., 2004; Oike et al., 1999; Wood et al., 2005; Yin et al., 1994) (but see Balschun et al., 2003).
A third pathway, involves signaling by target of rapamycin (TOR) proteins. The TORs are evolutionarily conserved protein kinases that regulate the balance between protein synthesis and degradation. Kandel and coworkers implicated TOR signaling in long-term facilitation in Apylsia neurons by demonstrating that serotonin stimulated synaptic protein synthesis can be blocked with rapamycin, a potent inhibitor of TOR proteins (Casadio et al., 1999). In mammals TOR or mTOR is thought to modulate translation of mRNAs via the regulation of the phosphorylation state of several different translation effector proteins including the ribosomal S6 kinases, 4E-BPs, eIF4GI, eEF2, and eIF4B although much is unknown of these mechanisms (Carroll, Dyer, & Sossin, 2006; Raught, Gingras, & Sonenberg, 2001).
Knowledge of the intricacies of these ubiquitous signaling mechanisms (Deisseroth, Bito, & Tsien, 1996; Frank, 1994 #6638; Giovannini, 2006; Kaang, Kandel, & Grant, 1993; Lamprecht, Hazvi, & Dudai, 1997; Raught et al., 2001; Silva, Kogan, Frankland, & Kida, 1998; Warburton et al., 2005), although greatly informative on one hand, speaks little to the spatial regulation of translation within the neuron—an issue that created an interesting twist in our conception and definition of consolidation. Even so, we provide this summary not only to elaborate on consolidation theory but to demonstrate that despite the interpretational difficulties surrounding protein synthesis inhibitors great strides have been made that uphold the findings of studies using these “dirty” drugs. It could therefore be argued that molecular and genetic approaches have, in effect, negated worries over the non-specificity of protein synthesis inhibitors.
5. The changing role of protein synthesis in consolidation theory
The past few decades of research have shed light on activity-dependent processes that lead to changes in gene expression and the subsequent induction of protein synthesis in the soma. However, it is unclear how such changes are able to alter the strength of particular synaptic connections within a neuron. Returning to our previous question: could it be that de novo protein synthesis is not the main mechanism by which memories are permanently stabilized? Recent data suggests that the nucleus is not alone in having control over long-term memory formation but that local protein synthesis in the dendrites might also influence the synaptic stability required for the formation and maintenance of memory.
5.1. Stabilization by local protein synthesis?
The role of dendritic or “local” protein synthesis in learning and memory was first hypothesized in 1965, after Bodian demonstrated the existence of ribosomes in the dendrites of in monkey spinal cord motoneurons (Bodian, 1965). This was a significant finding in that it had been long-assumed that protein synthesis only took place in the soma. Ribosomes were subsequently found in the dendrites of neurons of dentate granule cell neurons (Steward & Levy, 1982). The incorporation of radiolabeled amino acids in proteins were subsequently detected in synaptic fractions (Rao & Steward, 1991; Torre & Steward, 1992; Weiler & Greenough, 1993) and in the dendrites of hippocampal slices with latencies to short to be the result of transcription and the subsequent transportation of newly synthesized protein from the soma to the dendrite (Feig & Lipton, 1993). Importantly, recent findings (reviewed by Sutton & Schuman, 2006) have demonstrated a role for local protein synthesis in the modulation of synaptic transmission (Kang & Schuman, 1996) offering a potential mechanism by which selective synaptic changes might be conferred to recently activated circuits. Thus, even with a wealth of data demonstrating the importance of transcription factors in the regulation of gene expression and protein synthesis during memory consolidation, we are now forced to reconsider the actual role these events play relative to those that occur at or near the synapse itself in response to learning-related stimuli.
5.2. Targeting by tagging
If local protein synthesis plays a significant role in memory formation it must be shown that a dialog between the nucleus and synapse exists whereby mRNAs are transported to those sites where local synthesis of proteins is needed. Indeed, specific mRNAs can be dendritically targeted (Garner, Tucker, & Matus, 1988; Martin, 2004; Martin & Zukin, 2006; Tongiorgi, Righi, & Cattaneo, 1997; Wang & Tiedge, 2004). For example, mouse CaMKIIα mRNA is dendritically localized via targeting sequences in its 3′UTR and undergoes activity-dependent synaptic translation (Mayford, Baranes, Podsypanina, & Kandel, 1996; Richter & Lorenz, 2002; Rook, Lu, & Kosik, 2000). Mice with a mutation in the 3′UTR region of the CaMKIIα gene display a reduction in CaMKIIα levels in post-synaptic densities and impaired L-LTP and long-term memory for a several tasks (Miller et al., 2002). Knowledge of the complete range of functional control afforded by local translation, however, is limited by the number of known mRNAs localized at the dendrite. How the products of gene expression might be efficiently targeted and transported to only the correct subset of synapses needed for memory formation within a particular neuron remains the subject of great interest and debate. The notion of synaptic “tagging” or “capture” has been proposed to deal with such a problem (Reymann & Frey, 2007). Using LTP as a model for memory, Frey and Morris originally proposed that a temporary protein synthesis-independent tag is established at potentiated synapses that sequesters the relevant proteins needed to maintain LTP. In support of this notion, they found that weak stimulation of one set of synapses, which ordinarily leads to E-LTP, surprisingly resulted in the expression of L-LTP given the strong stimulation of a separate set of synapses terminating on the same post-synaptic neuron had just occurred (Frey & Morris, 1997; Frey & Morris, 1998).
Studies conducted by Martin and colleagues in Aplysia have also provided evidence of tagging or capture. In a culture system consisting of a single bifurcated sensory neuron that had formed synaptic connections with two spatially separated motor neurons, a series of five pulses of serotonin (5-HT) applied to one synapse resulted in facilitation of that synapse, whereas the potency of the other was unchanged. Importantly, this effect was transcription and CREB dependent demonstrating that the nucleus had some role in directing plasticity to a specific, previously activated synapse (Martin et al., 1997). Casadio and colleagues then demonstrated that the application of five pulses of 5-HT to the soma of the sensory neuron lead to a temporary, protein synthesis-dependent, cell wide form of facilitation that could be captured by specific synapses after administration of one additional pulse of 5-HT. Thus, both instructive protein synthesis mediated by the nucleus in addition to local protein synthesis at the synapse appear to play a role in memory storage (Casadio et al., 1999). To explain these results the tagging hypothesis proposes that gene products are dispersed throughout the cell but that they function to increase synaptic strength only in activated synapses that have recently been tagged. The identity of the tag, although still unknown, might be a single molecule, a complex of molecules, or a process such as activation of local translation, cytoskeletal reorganization, or retrograde signaling (Martin & Kosik, 2002). Importantly, no behavioral correlate of synaptic tagging has been demonstrated and therefore tagging remains a controversial matter.
Regardless, its seems that consolidation theory must be modified to account for the ability of dendrites to regulate protein synthesis at the level of the synapse, at least to some extent, without activity-dependent instructions originating in the nucleus. However, such a modification to consolidation theory has important implications. For example, what percentage of learning- or activity-dependent protein synthesis serves to simply replenish stocks of proteins or mRNAs at the synapse or dendrite? How “instructive” or involved is the genetic program of the neuron at the time of memory formation? To what extent is this program initiated before learning-related events occur (i.e., during the development of the neuron)? Is gene expression involved primarily in the maintenance of memory? All of these questions speak to whether memory lies at the synapse or in the nucleus. It is clear, however, that tactics beyond the use of protein synthesis inhibitors will be required to answer these important questions.
6. New approaches in the investigation of protein synthesis in memory formation
The inadequacies of protein synthesis inhibitors still in common use must be overcome and new methods that target translation must be developed. Indeed, the challenge we currently face in terms understanding learning-related protein synthesis is to demonstrate that which is induced not what is needed for general housekeeping functions. Here we summarize how targeted disruption of translation through genetic and pharmaco-genetic approaches has recently been employed to explore the role of protein synthesis in memory consolidation without depending on protein synthesis inhibitors. One interesting genetic approach utilized mice that lacked the protein kinase GCN2, a negative regulator of translation initiation that acts by phosphorylating eukaryotic initiation factor 2α (eIF2α). It was shown that the threshold for eliciting L-LTP in hippocampal slices from these mice was lowered and spatial memory was enhanced after weak training in the Morris water maze (Costa-Mattioli et al., 2005). In a similar approach, mice were engineered that lacked the translation repressor eukaryotic initiation factor 4E-binding protein 2 (4E-BP2). Interestingly, these mice demonstrated memory enhancements and impairments depending on the task (Banko, Hou, Poulin, Sonenberg, & Klann, 2006). Thus, it is possible to disrupt translation without the use of nonspecific pharmacological agents.
In a clever experiment reported by Sonenberg and colleagues, the phenotype of mice with a mutated form of eIF2α (eIF2α(+/S51A) mice) was reversed by a simple pharmacological treatment. eIF2α(+/S51A) mice, in which eIF2α phosphorylation is reduced, display enhanced LTP and enhanced performance in several tests of memory, however, upon administration of a small molecule inhibitor of eIF2α dephosphorylation, Sal003, L-LTP and memory were blocked (Costa-Mattioli et al., 2007). Importantly, these types of studies demonstrate not only the role of specific proteins in memory formation but also the power in combining genetically encoded methods of protein synthesis inhibition with relatively temporally precise methods of pharmacological control. Further development of these systems through the use of specific promoters to alter gene/protein function within specific regions of the brain or neuron will further enhance their utility.
Behavioral assays of pharmacological and genetic manipulations, as informative as they may be, are limited in that it is impossible to see events occurring in “real time”. Cellular imaging techniques making use of fluorescent reporters to monitor changes in translation (Aakalu, Smith, Nguyen, Jiang, & Schuman, 2001; Ashraf, McLoon, Sclarsic, & Kunes, 2006; Martin & Zukin, 2006; Reijmers, Perkins, Matsuo, & Mayford, 2007) are revealing much about how plasticity-related translation is temporally and spatially regulated within the neuron. For example, imaging techniques have been used to assess the effects of interference with the normal dendritic targeting of specific mRNAs (Miller et al., 2002) and to investigate the means by which translation of mRNA remains repressed in the absence of neuronal activity (Klann & Dever, 2004; Liu, Valencia-Sanchez, Hannon, & Parker, 2005). Additionally, real-time imaging of fluorescent reporter proteins after the induction of LTP together with pharmacological manipulations has led to the novel finding that LTP induction in the hippocampus leads not only to an increase in the rate of protein synthesis but also to an increase in the active degradation of proteins (Karpova, Mikhaylova, Thomas, Knopfel, & Behnisch, 2006). Imaging can also be used to begin to investigate which proteins are synthesized at the synapse and which are transported there and by what mechanism. In fact, these techniques could be used to image gene expression itself. To this end, identification of the set of mRNAs present in dendrites will help elucidate how protein synthesis might be controlled and the role it plays in memory consolidation (Miyashiro, Dichter, & Eberwine, 1994; Moccia et al., 2003; Zhong, Zhang, & Bloch, 2006).
7. Alternatives to consolidation theory
7.1. Post-translational modification of synaptic proteins in memory formation
In light of interpretational difficulties regarding the effects protein synthesis inhibitors on memory (and a lack of pharmacological agents that are able to specifically inhibit protein synthesis), Routtenberg and Rekart have proposed a provocative, less “translation-centric” model of memory formation centering, instead, on the function of post-translational modifications (PTMs) to the local milieu of pre-existing synaptic proteins (Routtenberg & Rekart, 2005).
At its core, PTM theory, at least as we interpret it, proposes that activity-dependent changes in the PTM state of synaptic proteins are sufficient to control various processes including the structural remodeling of synapses, changes in synaptic strength, and intracellular signaling amongst other cellular changes needed for long-term memory formation. Further, PTM theory posits that all transcription and translation occurs solely to replenish the molecules consumed in the consolidation process, in contrast to the more instructive role of the gene expression endorsed by consolidation theory. Appropriate levels of synaptic proteins would be maintained by concentration gradients and through retrograde signaling mechanisms from the post-synaptic to pre-synaptic neurons composing a sort of synaptic “dialogue” throughout the network in which the memory trace is located. Although an in-depth critique of PTM theory lies outside the scope of this review, the reader is encouraged to reference the original article for a more complete description and to evaluate whether the mechanisms cited in support of PTM necessarily exclude those put forth by consolidation theory or whether some composite of both theories might best guide future research.
However, we do consider some preliminary work aimed at testing the validity of PTM theory as the conclusions drawn from these data embody the type interpretational difficulties highlighted in this review. As an initial test of PTM theory, Holahan and Routtenberg examined the effect of alterations to the phosphorylation state of several molecules (e.g., PKA, PKC, PKG) on long-term memory through intracerebral infusions of the broad spectrum serine/threonine inhibitor [1-(5-isoquinoline-sulphonyl)-2-methylpiperazine] (H7) (Holahan & Routtenberg, 2007). The authors predicted that if the PTM states of synaptic proteins play an enduring role in the maintenance of memory then even remote memories should be sensitive to disruption by alterations in the PTM state of critical proteins. In rats, they demonstrate that a single infusion of H7 aimed at the anterior cingulate 1 h prior to testing impaired freezing 3 weeks after contextual fear conditioning. Furthermore, in mice, they report that intra-medial prefrontal cortex infusions of H7 impaired extinction of the freezing response in a similar paradigm. The observed impairments were taken as initial evidence in support of PTM theory. However, there are plausible alternative interpretations of their results. For example, pre-testing infusions of H7 could have simply impaired retrieval of contextual fear memory in the first experiment. In the second experiment, pre-training infusions of H7 could have interrupted the encoding or consolidation of the extinction memory by disrupting PKA-dependent translation events. Thus, it should be reemphasized that the use of semi-selective pharmacological agents are not sufficient to isolate the function of various cellular processes, in this case, to delineate the function of PTM states of translation-related signaling kinases from protein synthesis on memory formation.
Another aspect of PTM theory questions the role of local protein synthesis because proteins produced in the soma are indistinguishable from dendritically translated proteins and that some mRNAs for plasticity-related proteins have not yet been found in dendrites. However, we feel these are simple technical limitations and that techniques utilizing live imaging of tagged proteins may soon be able to provide a clear picture of the spatial and temporal control of protein synthesis. Routtenberg and Rekart also reject synaptic tagging as a means of directing replenishment because, because tagging does not require the type of ongoing synaptic dialog they propose is necessary to maintain the trace over long periods of time. However, recent studies by Schuman and colleagues have begun to address how a form of synaptic dialogue carried out through the frequency of spontaneous miniature excitatory postsynaptic currents (minis) might influence the ongoing organization of the synapse. They reported that inhibition of minis leads to the enhancement of dendritic protein synthesis. Additionally, blockade of the NMDA receptor component of minis rapidly increased the compensatory responsiveness to glutamate at the synapse in a protein synthesis dependent fashion (Sutton & Schuman, 2006; Sutton, Wall, Aakalu, & Schuman, 2004). The authors conclude that minis (viewed here as a component of synaptic dialogue) helped maintain synaptic stability by tonically repressing the dendritic protein-synthesis machinery. Thus, there may be an important link between minis (synaptic dialogue), the maintenance of the PTM state of synaptic proteins, local protein synthesis, and tagging.
7.2. Sleep to consolidate: A role for sleep-induced protein synthesis in memory
A growing body of literature suggests that changes in the electrophysiological state of the brain during sleep play an important role in the consolidation and organization of memory potentially through protein synthesis-dependent mechanisms. This notion is supported by several lines of evidence. First sleep deprivation is known to impair memory and LTP (Campbell, Guinan, & Horowitz, 2002; Davis, Harding, & Wright, 2003; Gais & Born, 2004; Graves, Heller, Pack, & Abel, 2003; Kim, Mahmoud, & Grover, 2005). Indeed, both of the major types of sleep, rapid eye movement (REM) and non-rapid eye movement (NREM) sleep, are altered by learning and each have been shown to influence performance in a variety of memory tasks (Best, Diniz Behn, Poe, & Booth, 2007; Buzsaki, 1989; Hellman & Abel, 2007; Stickgold & Walker, 2007). Electrophysiological data suggest that reactivation of specific neurons within cortico-thalamic networks after learning or other wakeful experiences serves as a rehersal mechanism whereby NMDA receptor reactivation might mediate further strengthening of previously consolidated memory via protein synthesis-dependent mechanisms (Datta, Mavanji, Ulloor, & Patterson, 2004; Kudrimoti, Barnes, & McNaughton, 1999; Lee & Wilson, 2002; Pavlides & Winson, 1989; Wang, Hu, & Tsien, 2006; Wilson & McNaughton, 1994). Importantly, DNA microarray studies and proteomic studies have demonstrated changes in plasticity-related gene expression during sleep (Basheer, Brown, Ramesh, Begum, & McCarley, 2005; Cirelli, Faraguna, & Tononi, 2006; Cirelli, LaVaute, & Tononi, 2005; Cirelli & Tononi, 1999; Mackiewicz et al., 2007). Finally, there is ample evidence that behavioral readouts of memory are significantly enhanced by sleep (reviewed by Walker, 2005).
Thus, it may be that the electrophysiological state of the brain during sleep recapitulates the activity in neural circuits that occurred during previous bouts of wakefulness (Ji & Wilson, 2007) and in doing so might initiate signal transduction cascades similar to those initiated by the original learning experience thereby enhancing previously consolidated memory (Born, Rasch, & Gais, 2006; Ganguly-Fitzgerald, Donlea, & Shaw, 2006). Nevertheless, the functionality of sleep and the molecular mechanisms by which sleep improves memory remains largely unknown (Frank, 2006; Molter, Sato, & Yamaguchi, 2007) (but see O’Hara, Ding, Bernat, & Franken, 2007; Ribeiro et al., 2002). However, it is clear that a deeper understanding of the nature of gene expression and translation that occur during sleep is required to truly understand the nature of memory.
8. Final thoughts
It would be difficult to identify a scientific field that has benefited so dramatically from the contributions of such numerous and diverse methodologies as has the study of learning and memory. Approaches deriving from behavior, electrophysiology, pharmacology, genetics and molecular and cellular biology have been indispensable tools in unlocking the riddles of memory formation. Although there may not be agreement on the nature of the molecular mechanisms underlying amnesia induced by protein synthesis inhibitors, these drugs have played a pivotal role in our understanding of memory formation from the role of NMDA and other receptors in initiating intracellular kinase signaling cascades, to the ability of these cascades to direct new protein synthesis and gene expression. However, we have come to a stalemate in the debate over the precise role of protein synthesis in memory consolidation. Indeed, while protein synthesis inhibitors have been a useful tool, our questions are becoming more sophisticated where the specificity of these pharmacological tools is not.
A major theme in this review is whether protein synthesis plays an instructive or permissive role in memory consolidation. Is memory actually stored in the nucleus as a set of epigenetic marks or does the synapse/dendrite have everything it needs such that translation functions to simply replenish proteins consumed in the consolidation process? If the latter case is true, how does efficient replenishment occur? Is synaptic tagging a behaviorally relevant method of marking previous synaptic activity? Are both permissive and instructive processes at work simultaneously?
Indeed, we are at a time when definitions are being rewritten as nuances in memory processing are being elucidated. What is referred to as “consolidation” no doubt encompasses what will someday be parceled more precisely into separate processes. However, the challenges facing learning and memory research can no longer be addressed through the use of non-specific protein synthesis inhibitors and other pharmacological agents. By devising methods of altering the function of proteins that are induced by learning and by developing ways to image processes induced by learning (Reijmers et al., 2007), we will begin to provide answers to these questions. Yet, we must recognize that it is largely because of a handful of protein synthesis inhibitors that we are closer to answering the ultimate question: what is memory?
Footnotes
Publisher's Disclaimer: This article was published in an Elsevier journal. The attached copy is furnished to the author for non-commercial research and education use, including for instruction at the author's institution, sharing with colleagues and providing to institution administration.
Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited.
In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier's archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright
References
- Aakalu G, Smith WB, Nguyen N, Jiang C, Schuman EM. Dynamic visualization of local protein synthesis in hippocampal neurons. Neuron. 2001;30:489–502. doi: 10.1016/s0896-6273(01)00295-1. [DOI] [PubMed] [Google Scholar]
- Abel T, Lattal KM. Molecular mechanisms of memory acquisition, consolidation and retrieval. Current Opinion in Neurobiology. 2001;11:180–187. doi: 10.1016/s0959-4388(00)00194-x. [DOI] [PubMed] [Google Scholar]
- Abel T, Nguyen PV, Barad M, Deuel TA, Kandel ER, Bourtchouladze R. Genetic demonstration of a role for PKA in the late phase of LTP and in hippocampus-based long-term memory. Cell. 1997;88:615–626. doi: 10.1016/s0092-8674(00)81904-2. [DOI] [PubMed] [Google Scholar]
- Abraham WC, Mason SE, Demmer J, Williams JM, Richardson CL, Tate WP, et al. Correlations between immediate early gene induction and the persistence of long-term potentiation. Neuroscience. 1993;56:717–727. doi: 10.1016/0306-4522(93)90369-q. [DOI] [PubMed] [Google Scholar]
- Acsady L, Kali S. Models, structure, function: The transformation of cortical signals in the dentate gyrus. Progress in Brain Research. 2007;163:577–599. doi: 10.1016/S0079-6123(07)63031-3. [DOI] [PubMed] [Google Scholar]
- Agranoff BW. Memory and protein synthesis. Scientific American. 1967;216:115–122. doi: 10.1038/scientificamerican0667-115. [DOI] [PubMed] [Google Scholar]
- Agranoff BW, Davis RE, Brink JJ. Memory fixation in the goldfish. Proceedings of the National Academy of Sciences of the United States of America. 1965;54:788–793. doi: 10.1073/pnas.54.3.788. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Agranoff BW, Davis RE, Casola L, Lim R. Actinomycin D blocks formation of memory of shock-avoidance in goldfish. Science. 1967;158:1600–1601. doi: 10.1126/science.158.3808.1600. [DOI] [PubMed] [Google Scholar]
- Agranoff BW, Klinger PD. Puromycin effect on memory fixation in the goldfish. Science. 1964;146:952–953. doi: 10.1126/science.146.3646.952. [DOI] [PubMed] [Google Scholar]
- Alberini CM. Mechanisms of memory stabilization: Are consolidation and reconsolidation similar or distinct processes? Trends in Neurosciences. 2005;28:51–56. doi: 10.1016/j.tins.2004.11.001. [DOI] [PubMed] [Google Scholar]
- Alger BE, Teyler TJ. Long-term and short-term plasticity in the CA1, CA3, and dentate regions of the rat hippocampal slice. Brain Research. 1976;110:463–480. doi: 10.1016/0006-8993(76)90858-1. [DOI] [PubMed] [Google Scholar]
- Allweis C. The congruity of rat and chick multiphasic memory-consolidation models. In: Andrew RJ, editor. Neural and behavioural plasticity. New York: Oxford University Press; 1991. pp. 370–393. [Google Scholar]
- Apergis-Schoute AM, Debiec J, Doyere V, LeDoux JE, Schafe GE. Auditory fear conditioning and long-term potentiation in the lateral amygdala require ERK/MAP kinase signaling in the auditory thalamus: A role for presynaptic plasticity in the fear system. Journal of Neuroscience. 2005;25:5730–5739. doi: 10.1523/JNEUROSCI.0096-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ashraf SI, McLoon AL, Sclarsic SM, Kunes S. Synaptic protein synthesis associated with memory is regulated by the RISC pathway in Drosophila. Cell. 2006;124:191–205. doi: 10.1016/j.cell.2005.12.017. [DOI] [PubMed] [Google Scholar]
- Bacskai BJ, Hochner B, Mahaut-Smith M, Adams SR, Kaang BK, Kandel ER, et al. Spatially resolved dynamics of cAMP and protein kinase A subunits in Aplysia sensory neurons. Science. 1993;260:222–226. doi: 10.1126/science.7682336. [DOI] [PubMed] [Google Scholar]
- Bailey CH, Bartsch D, Kandel ER. Toward a molecular definition of long-term memory storage. Proceedings of the National Academy of Sciences of the United States of America. 1996;93:13445–13452. doi: 10.1073/pnas.93.24.13445. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bailey DJ, Kim JJ, Sun W, Thompson RF, Helmstetter FJ. Acquisition of fear conditioning in rats requires the synthesis of mRNA in the amygdala. Behavioral Neuroscience. 1999;113:276–282. doi: 10.1037//0735-7044.113.2.276. [DOI] [PubMed] [Google Scholar]
- Balschun D, Wolfer DP, Gass P, Mantamadiotis T, Welzl H, Schutz G, et al. Does cAMP response element-binding protein have a pivotal role in hippocampal synaptic plasticity and hippocampus-dependent memory? Journal of Neuroscience. 2003;23:6304–6314. doi: 10.1523/JNEUROSCI.23-15-06304.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Banko JL, Hou L, Poulin F, Sonenberg N, Klann E. Regulation of eukaryotic initiation factor 4E by converging signaling pathways during metabotropic glutamate receptor-dependent long-term depression. Journal of Neuroscience. 2006;26:2167–2173. doi: 10.1523/JNEUROSCI.5196-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barnes CA. Memory deficits associated with senescence: A neurophysiological and behavioral study in the rat. Journal of Comparative and Physiological Psychology. 1979;93:74–104. doi: 10.1037/h0077579. [DOI] [PubMed] [Google Scholar]
- Barondes SH, Cohen HD. Puromycin effect on successive phases of memory storage. Science. 1966;151:594–595. doi: 10.1126/science.151.3710.594. [DOI] [PubMed] [Google Scholar]
- Barondes SH, Cohen HD. Comparative effects of cycloheximide and puromycin on cerebral protein synthesis and consolidation of memory in mice. Brain Research. 1967;4:44–51. doi: 10.1016/0006-8993(67)90147-3. [DOI] [PubMed] [Google Scholar]
- Barondes SH, Jarvik ME. The Influence of Actinomycin-D on Brain Rna Synthesis and on Memory. Journal of Neurochemistry. 1964;11:187–195. doi: 10.1111/j.1471-4159.1964.tb06128.x. [DOI] [PubMed] [Google Scholar]
- Barzilai A, Kennedy TE, Sweatt JD, Kandel ER. 5-HT modulates protein synthesis and the expression of specific proteins during long-term facilitation in Aplysia sensory neurons. Neuron. 1989;2:1577–1586. doi: 10.1016/0896-6273(89)90046-9. [DOI] [PubMed] [Google Scholar]
- Basheer R, Brown R, Ramesh V, Begum S, McCarley RW. Sleep deprivation-induced protein changes in basal forebrain: Implications for synaptic plasticity. Journal of Neuroscience Research. 2005;82:650–658. doi: 10.1002/jnr.20675. [DOI] [PubMed] [Google Scholar]
- Bekinschtein P, Cammarota M, Igaz LM, Bevilaqua LR, Izquierdo I, Medina JH. Persistence of long-term memory storage requires a late protein synthesis- and BDNF-dependent phase in the hippocampus. Neuron. 2007;53:261–277. doi: 10.1016/j.neuron.2006.11.025. [DOI] [PubMed] [Google Scholar]
- Berman DE, Dudai Y. Memory extinction, learning anew, and learning the new: Dissociations in the molecular machinery of learning in cortex. Science. 2001;291:2417–2419. doi: 10.1126/science.1058165. [DOI] [PubMed] [Google Scholar]
- Best J, Diniz Behn C, Poe GR, Booth V. Neuronal models for sleep–wake regulation and synaptic reorganization in the sleeping hippocampus. Journal of Biological Rhythms. 2007;22:220–232. doi: 10.1177/0748730407301239. [DOI] [PubMed] [Google Scholar]
- Bliss TV, Collingridge GL. A synaptic model of memory: Long-term potentiation in the hippocampus. Nature. 1993;361:31–39. doi: 10.1038/361031a0. [DOI] [PubMed] [Google Scholar]
- Bliss TV, Lømo T. Long-lasting potentiation of synaptic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. J Physiol. 1973;232:331–356. doi: 10.1113/jphysiol.1973.sp010273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blokland A, Schreiber R, Prickaerts J. Improving memory: A role for phosphodiesterases. Current Pharmaceutical Design. 2006;12:2511–2523. doi: 10.2174/138161206777698855. [DOI] [PubMed] [Google Scholar]
- Bodian D. A Suggestive Relationship of Nerve Cell Rna with Specific Synaptic Sites. Proceedings of the National Academy of Sciences of the United States of America. 1965;53:418–425. doi: 10.1073/pnas.53.2.418. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Born J, Rasch B, Gais S. Sleep to remember. Neuroscientist. 2006;12:410–424. doi: 10.1177/1073858406292647. [DOI] [PubMed] [Google Scholar]
- Bourtchouladze R, Abel T, Berman N, Gordon R, Lapidus K, Kandel ER. Different training procedures recruit either one or two critical periods for contextual memory consolidation, each of which requires protein synthesis and PKA. Learning and Memory. 1998;5:365–374. [PMC free article] [PubMed] [Google Scholar]
- Bourtchouladze R, Frenguelli B, Blendy J, Cioffi D, Schutz G, Silva AJ. Deficient long-term memory in mice with a targeted mutation of the cAMP-responsive element-binding protein. Cell. 1994;79:59–68. doi: 10.1016/0092-8674(94)90400-6. [DOI] [PubMed] [Google Scholar]
- Bourtchouladze R, Lidge R, Catapano R, Stanley J, Gossweiler S, Romashko D, et al. A mouse model of Rubinstein–Taybi syndrome: Defective long-term memory is ameliorated by inhibitors of phosphodiesterase 4. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:10518–10522. doi: 10.1073/pnas.1834280100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brandon EP, Zhuo M, Huang YY, Qi M, Gerhold KA, Burton KA, et al. Hippocampal long-term depression and depotentiation are defective in mice carrying a targeted disruption of the gene encoding the RI beta subunit of cAMP-dependent protein kinase. Proceedings of the National Academy of Sciences of the United States of America. 1995;92:8851–8855. doi: 10.1073/pnas.92.19.8851. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Buzsaki G. Two-stage model of memory trace formation: A role for “noisy” brain states. Neuroscience. 1989;31:551–570. doi: 10.1016/0306-4522(89)90423-5. [DOI] [PubMed] [Google Scholar]
- Byers D, Davis RL, Kiger JA., Jr Defect in cyclic AMP phosphodiesterase due to the dunce mutation of learning in Drosophila melanogaster. Nature. 1981;289:79–81. doi: 10.1038/289079a0. [DOI] [PubMed] [Google Scholar]
- Campbell IG, Guinan MJ, Horowitz JM. Sleep deprivation impairs long-term potentiation in rat hippocampal slices. Journal of Neurophysiology. 2002;88:1073–1076. doi: 10.1152/jn.2002.88.2.1073. [DOI] [PubMed] [Google Scholar]
- Carroll M, Dyer J, Sossin WS. Serotonin increases phosphorylation of synaptic 4EBP through TOR, but eukaryotic initiation factor 4E levels do not limit somatic cap-dependent translation in aplysia neurons. Molecular and Cellular Biology. 2006;26:8586–8598. doi: 10.1128/MCB.00955-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Casadio A, Martin KC, Giustetto M, Zhu H, Chen M, Bartsch D, et al. A transient, neuron-wide form of CREB-mediated long-term facilitation can be stabilized at specific synapses by local protein synthesis. Cell. 1999;99:221–237. doi: 10.1016/s0092-8674(00)81653-0. [DOI] [PubMed] [Google Scholar]
- Castellucci VF, Frost WN, Goelet P, Montarolo PG, Schacher S, Morgan JA, et al. Cell and molecular analysis of long-term sensitization in Aplysia. Journal de Physiologie. 1986;81:349–357. [PubMed] [Google Scholar]
- Chang L, Karin M. Mammalian MAP kinase signalling cascades. Nature. 2001;410:37–40. doi: 10.1038/35065000. [DOI] [PubMed] [Google Scholar]
- Cherkin A. Kinetics of memory consolidation: Role of amnesic treatment parameters. Proceedings of the National Academy of Sciences of the United States of America. 1969;63:1094–1101. doi: 10.1073/pnas.63.4.1094. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chew SJ, Vicario DS, Nottebohm F. Quantal duration of auditory memories. Science. 1996;274:1909–1914. doi: 10.1126/science.274.5294.1909. [DOI] [PubMed] [Google Scholar]
- Cirelli C, Faraguna U, Tononi G. Changes in brain gene expression after long-term sleep deprivation. Journal of Neurochemistry. 2006;98:1632–1645. doi: 10.1111/j.1471-4159.2006.04058.x. [DOI] [PubMed] [Google Scholar]
- Cirelli C, LaVaute TM, Tononi G. Sleep and wakefulness modulate gene expression in Drosophila. Journal of Neurochemistry. 2005;94:1411–1419. doi: 10.1111/j.1471-4159.2005.03291.x. [DOI] [PubMed] [Google Scholar]
- Cirelli C, Tononi G. Differences in brain gene expression between sleep and waking as revealed by mRNA differential display and cDNA microarray technology. Journal of Sleep Research. 1999;8(Suppl 1):44–52. doi: 10.1046/j.1365-2869.1999.00008.x. [DOI] [PubMed] [Google Scholar]
- Codish SD. Actinomycin D injected into the hippocampus of chicks: Effects upon imprinting. Physiology and Behavior. 1971;6:95–96. doi: 10.1016/0031-9384(71)90024-2. [DOI] [PubMed] [Google Scholar]
- Coffey ET, Hongisto V, Dickens M, Davis RJ, Courtney MJ. Dual roles for c-Jun N-terminal kinase in developmental and stress responses in cerebellar granule neurons. Journal of Neuroscience. 2000;20:7602–7613. doi: 10.1523/JNEUROSCI.20-20-07602.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Collingridge GL, Kehl SJ, McLennan H. Excitatory amino acids in synaptic transmission in the Schaffer collateral-commissural pathway of the rat hippocampus. J Physiol. 1983;334:33–46. doi: 10.1113/jphysiol.1983.sp014478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Costa-Mattioli M, Gobert D, Harding H, Herdy B, Azzi M, Bruno M, et al. Translational control of hippocampal synaptic plasticity and memory by the eIF2alpha kinase GCN2. Nature. 2005;436:1166–1173. doi: 10.1038/nature03897. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Costa-Mattioli M, Gobert D, Stern E, Gamache K, Colina R, Cuello C, et al. eIF2alpha phosphorylation bidirectionally regulates the switch from short- to long-term synaptic plasticity and memory. Cell. 2007;129:195–206. doi: 10.1016/j.cell.2007.01.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cotman CW, Bridges RJ, Taube JS, Clark AS, Geddes JW, Monaghan DT. The role of the NMDA receptor in central nervous system plasticity and pathology. Journal of the National Institutes of Health Research. 1989;1:65–74. [Google Scholar]
- Crow T, Siddiqi V, Dash PK. Long-term enhancement but not short-term in Hermissenda is dependent upon mRNA synthesis. Neurobiology of Learning and Memory. 1997;68:343–350. doi: 10.1006/nlme.1997.3779. [DOI] [PubMed] [Google Scholar]
- Dash PK, Hochner B, Kandel ER. Injection of the cAMP-responsive element into the nucleus of Aplysia sensory neurons blocks long-term facilitation. Nature. 1990;345:718–721. doi: 10.1038/345718a0. [DOI] [PubMed] [Google Scholar]
- Dash PK, Moore AN, Kobori N, Runyan JD. Molecular activity underlying working memory. Learning and Memory. 2007;14:554–563. doi: 10.1101/lm.558707. [DOI] [PubMed] [Google Scholar]
- Datta S, Mavanji V, Ulloor J, Patterson EH. Activation of phasic pontine-wave generator prevents rapid eye movement sleep deprivation-induced learning impairment in the rat: A mechanism for sleep-dependent plasticity. Journal of Neuroscience. 2004;24:1416–1427. doi: 10.1523/JNEUROSCI.4111-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Davis CJ, Harding JW, Wright JW. REM sleep deprivation-induced deficits in the latency-to-peak induction and maintenance of long-term potentiation within the CA1 region of the hippocampus. Brain Research. 2003;973:293–297. doi: 10.1016/s0006-8993(03)02508-3. [DOI] [PubMed] [Google Scholar]
- Davis S, Laroche S. Mitogen-activated protein kinase/extracellular regulated kinase signalling and memory stabilization: A review. Genes, Brain and Behaviour. 2006;5(Suppl 2):61–72. doi: 10.1111/j.1601-183X.2006.00230.x. [DOI] [PubMed] [Google Scholar]
- Davis HP, Squire LR. Protein synthesis and memory: A review. Psychological Bulletin. 1984;96:518–559. [PubMed] [Google Scholar]
- Deisseroth K, Bito H, Tsien RW. Signaling from synapse to nucleus: Postsynaptic CREB phosphorylation during multiple forms of hippocampal synaptic plasticity. Neuron. 1996;16:89–101. doi: 10.1016/s0896-6273(00)80026-4. [DOI] [PubMed] [Google Scholar]
- DeZazzo J, Tully T. Dissection of memory formation: From behavioral pharmacology to molecular genetics. Trends in Neurosciences. 1995;18:212–218. doi: 10.1016/0166-2236(95)93905-d. [DOI] [PubMed] [Google Scholar]
- Doty BA, Doty LA. Facilitative effects of amphetamine on avoidance conditioning in relation to age and problem difficulty. Psychopharmacologia. 1966;9:234–241. doi: 10.1007/BF02198483. [DOI] [PubMed] [Google Scholar]
- Dubnau J, Tully T. Gene discovery in Drosophila: New insights for learning and memory. Annual Review of Neuroscience. 1998;21:407–444. doi: 10.1146/annurev.neuro.21.1.407. [DOI] [PubMed] [Google Scholar]
- Dudai Y. The neurobiology of consolidations, or, how stable is the engram? Annual Review of Psychology. 2004;55:51–86. doi: 10.1146/annurev.psych.55.090902.142050. [DOI] [PubMed] [Google Scholar]
- Dudai Y. Reconsolidation: The advantage of being refocused. Current Opinion in Neurobiology. 2006;16:174–178. doi: 10.1016/j.conb.2006.03.010. [DOI] [PubMed] [Google Scholar]
- Dudai Y, Eisenberg M. Rites of passage of the engram: Reconsolidation and the lingering consolidation hypothesis. Neuron. 2004;44:93–100. doi: 10.1016/j.neuron.2004.09.003. [DOI] [PubMed] [Google Scholar]
- Dudai Y, Jan YN, Byers D, Quinn WG, Benzer S. dunce, a mutant of Drosophila deficient in learning. Proceedings of the National Academy of Sciences of the United States of America. 1976;73:1684–1688. doi: 10.1073/pnas.73.5.1684. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duncan CP. The retroactive effect of electroshock on learning. Journal Of Comparative And Physiological Psychology. 1949;42:32–44. doi: 10.1037/h0058173. [DOI] [PubMed] [Google Scholar]
- Dunn AJ, Gray HE, Iuvone PM. Protein synthesis and amnesia: Studies with emetine and pactamycin. Pharmacology, Biochemistry and Behavior. 1977;6:1–4. doi: 10.1016/0091-3057(77)90151-4. [DOI] [PubMed] [Google Scholar]
- Ebbinghaus H. Über das Gedchtnis. Untersuchungen zur experimentellen Psychologie. Leipzig: Duncker and Humblot; 1885. [Google Scholar]
- Eccles JC. Hippocampal plasticity. Progress in Brain Research. 1979;51:133–138. doi: 10.1016/S0079-6123(08)61301-1. [DOI] [PubMed] [Google Scholar]
- Eisenberg M, Kobilo T, Berman DE, Dudai Y. Stability of retrieved memory: Inverse correlation with trace dominance. Science. 2003;301:1102–1104. doi: 10.1126/science.1086881. [DOI] [PubMed] [Google Scholar]
- Eliot LS, Dudai Y, Kandel ER, Abrams TW. Ca2+/calmodulin sensitivity may be common to all forms of neural adenylate cyclase. Proceedings of the National Academy of Sciences of the United States of America. 1989;86:9564–9568. doi: 10.1073/pnas.86.23.9564. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eriksson M, Taskinen M, Leppa S. Mitogen activated protein kinase-dependent activation of c-Jun and c-Fos is required for neuronal differentiation but not for growth and stress response in PC12 cells. Journal of Cellular Physiology. 2007;210:538–548. doi: 10.1002/jcp.20907. [DOI] [PubMed] [Google Scholar]
- Fedulov V, Rex CS, Simmons DA, Palmer L, Gall CM, Lynch G. Evidence that long-term potentiation occurs within individual hippocampal synapses during learning. Journal of Neuroscience. 2007;27:8031–8039. doi: 10.1523/JNEUROSCI.2003-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feig S, Lipton P. Pairing the cholinergic agonist carbachol with patterned Schaffer collateral stimulation initiates protein synthesis in hippocampal CA1 pyramidal cell dendrites via a muscarinic, NMDA-dependent mechanism. Journal of Neuroscience. 1993;13:1010–1021. doi: 10.1523/JNEUROSCI.13-03-01010.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Finnegan KT, Karler R. Role for protein synthesis in the neurotoxic effects of methamphetamine in mice and rats. Brain Research. 1992;591:160–164. doi: 10.1016/0006-8993(92)90991-h. [DOI] [PubMed] [Google Scholar]
- Flexner LB, Flexner JB. Effect of acetoxycycloheximide and of an acetoxycycloheximide–puromycin mixture on cerebral protein synthesis and memory in mice. Proceedings of the National Academy of Sciences of the United States of America. 1966;55:369–374. doi: 10.1073/pnas.55.2.369. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flexner JB, Flexner LB. Restoration of expression of memory lost after treatment with puromycin. Proceedings of the National Academy of Sciences of the United States of America. 1967;57:1651–1654. doi: 10.1073/pnas.57.6.1651. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flexner LB, Flexner JB. Intracerebral saline: Effect on memory of trained mice treated with puromycin. Science. 1968a;159:330–331. doi: 10.1126/science.159.3812.330. [DOI] [PubMed] [Google Scholar]
- Flexner LB, Flexner JB. Studies on memory: The long survival of peptidyl-puromycin in mouse brain. Proceedings of the National Academy of Sciences of the United States of America. 1968b;60:923–927. doi: 10.1073/pnas.60.3.923. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flexner LB, Flexner JB, Roberts RB. Stages of memory in mice treated with acetoxycycloheximide before or immediately after learning. Proceedings of the National Academy of Sciences of the United States of America. 1966;56:730–735. doi: 10.1073/pnas.56.2.730. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flexner LB, Flexner JB, Stellar E. Memory in mice as affected by intracerebral puromycin. Science. 1963:57–59. doi: 10.1126/science.141.3575.57. [DOI] [PubMed] [Google Scholar]
- Flexner LB, Flexner JB, Stellar E. Memory and cerebral protein synthesis in mice as affected by graded amounts of puromycin. Experimental Neurology. 1965;13:264–272. doi: 10.1016/0014-4886(65)90114-7. [DOI] [PubMed] [Google Scholar]
- Flexner JB, Flexner LB, Stellar E, De La Haba G, Roberts RB. Inhibition of protein synthesis in brain and learning and memory following puromycin. Journal of Neurochemistry. 1962;9:595–605. doi: 10.1111/j.1471-4159.1962.tb04216.x. [DOI] [PubMed] [Google Scholar]
- Flexner LB, Goodman RH. Studies on memory: Inhibitors of protein synthesis also inhibit catecholamine synthesis. Proceedings of the National Academy of Sciences of the United States of America. 1975;72:4660–4663. doi: 10.1073/pnas.72.11.4660. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flexner LB, Serota RG, Goodman RH. Cycloheximide and acetoxycycloheximide: Inhibition of tyrosine hydroxylase activity and amnestic effects. Proceedings of the National Academy of Sciences of the United States of America. 1973;70:354–356. doi: 10.1073/pnas.70.2.354. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Flood JF, Bennett EL, Orme E, Rosenzweig MR. Relation of memory formation to controlled amounts of brain protein synthesis. Physiology and Behavior. 1975;15:97–102. doi: 10.1016/0031-9384(75)90285-1. [DOI] [PubMed] [Google Scholar]
- Flood JF, Cherkin A. Fluoxetine enhances memory processing in mice. Psychopharmacology. 1987;93:36–43. doi: 10.1007/BF02439584. [DOI] [PubMed] [Google Scholar]
- Flood JF, Rosenzweig MR, Bennett EL, Orme AE. The influence of duration of protein synthesis inhibition on memory. Physiology and Behavior. 1973;10:555–562. doi: 10.1016/0031-9384(73)90221-7. [DOI] [PubMed] [Google Scholar]
- Flood JF, Smith GE, Bennett EL, Alberti MH, Orme AE, Jarvik ME. Neurochemical and behavioral effects of 306 P.J. Hernandez, T. Abel/Neurobiology of Learning and Memory 89 (2008) 293–311 catecholamine and protein synthesis inhibitors in mice. Pharmacology, Biochemistry and Behavior. 1986;24:631–645. doi: 10.1016/0091-3057(86)90569-1. [DOI] [PubMed] [Google Scholar]
- Folkers E, Drain P, Quinn WG. Radish, a Drosophila mutant deficient in consolidated memory. Proceedings of the National Academy of Sciences of the United States of America. 1993;90:8123–8127. doi: 10.1073/pnas.90.17.8123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Folkers E, Waddell S, Quinn WG. The Drosophila radish gene encodes a protein required for anesthesia-resistant memory. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:17496–17500. doi: 10.1073/pnas.0608377103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Frank MG. The mystery of sleep function: Current perspectives and future directions. Reviews in the Neurosciences. 2006;17:375–392. doi: 10.1515/revneuro.2006.17.4.375. [DOI] [PubMed] [Google Scholar]
- Frankland PW, Bontempi B. The organization of recent and remote memories. Nat Rev Neurosci. 2005;6:119–130. doi: 10.1038/nrn1607. [DOI] [PubMed] [Google Scholar]
- Freeman FM, Rose SP, Scholey AB. Two time windows of anisomycin-induced amnesia for passive avoidance training in the day-old chick. Neurobiology of Learning and Memory. 1995;63:291–295. doi: 10.1006/nlme.1995.1034. [DOI] [PubMed] [Google Scholar]
- Frey U, Frey S, Schollmeier F, Krug M. Influence of actinomycin D, a RNA synthesis inhibitor, on long-term potentiation in rat hippocampal neurons in vivo and in vitro. J Physiol. 1996;490(Pt 3):703–711. doi: 10.1113/jphysiol.1996.sp021179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Frey U, Krug M, Reymann KG, Matthies H. Anisomycin, an inhibitor of protein synthesis, blocks late phases of LTP phenomena in the hippocampal CA1 region in vitro. Brain Research. 1988;452:57–65. doi: 10.1016/0006-8993(88)90008-x. [DOI] [PubMed] [Google Scholar]
- Frey U, Morris RG. Synaptic tagging and long-term potentiation. Nature. 1997;385:533–536. doi: 10.1038/385533a0. [DOI] [PubMed] [Google Scholar]
- Frey U, Morris RG. Weak before strong: Dissociating synaptic tagging and plasticity-factor accounts of late-LTP. Neuropharmacology. 1998;37:545–552. doi: 10.1016/s0028-3908(98)00040-9. [DOI] [PubMed] [Google Scholar]
- Frieder B, Allweis C. Transient hypoxic-amnesia: Evidence for a triphasic memory-consolidating mechanism with parallel processing. Behavioral Biology. 1978;22:178–189. doi: 10.1016/s0091-6773(78)92200-9. [DOI] [PubMed] [Google Scholar]
- Frieder B, Allweis C. Memory consolidation: Further evidence for the four-phase model from the time-courses of diethyldithiocarbamate and ethacrynic acid amnesias. Physiology and Behavior. 1982;29:1071–1075. doi: 10.1016/0031-9384(82)90300-6. [DOI] [PubMed] [Google Scholar]
- Gais S, Born J. Declarative memory consolidation: Mechanisms acting during human sleep. Learning and Memory. 2004;11:679–685. doi: 10.1101/lm.80504. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gale EF, Cundliffe E, Reynolds PE, Richmond MH, Waring MJ. The Molecular Basis of Antibiotic Action. New York: John Wiley & Sons Ltd; 1981. [Google Scholar]
- Ganguly-Fitzgerald I, Donlea J, Shaw PJ. Waking experience affects sleep need in Drosophila. Science. 2006;313:1775–1781. doi: 10.1126/science.1130408. [DOI] [PubMed] [Google Scholar]
- Gardner AM, Xu FH, Fady C, Jacoby FJ, Duffey DC, Tu Y, et al. Apoptotic vs. nonapoptotic cytotoxicity induced by hydrogen peroxide. Free Radical Biology and Medicine. 1997;22:73–83. doi: 10.1016/s0891-5849(96)00235-3. [DOI] [PubMed] [Google Scholar]
- Garner CC, Tucker RP, Matus A. Selective localization of messenger RNA for cytoskeletal protein MAP2 in dendrites. Nature. 1988;336:674–677. doi: 10.1038/336674a0. [DOI] [PubMed] [Google Scholar]
- Ghirardi M, Montarolo PG, Kandel ER. A novel intermediate stage in the transition between short- and long-term facilitation in the sensory to motor neuron synapse of aplysia. Neuron. 1995;14:413–420. doi: 10.1016/0896-6273(95)90297-x. [DOI] [PubMed] [Google Scholar]
- Gibbs ME, Ng KT. Behavioural stages in memory formation. Neuroscience Letters. 1979;13:279–283. doi: 10.1016/0304-3940(79)91507-6. [DOI] [PubMed] [Google Scholar]
- Giovannini MG. The role of the extracellular signal-regulated kinase pathway in memory encoding. Reviews in the Neurosciences. 2006;17:619–634. doi: 10.1515/revneuro.2006.17.6.619. [DOI] [PubMed] [Google Scholar]
- Glickman SE. Perseverative neural processes and consolidation of the memory trace. Psychological Bulletin. 1961;58:218–233. doi: 10.1037/h0044212. [DOI] [PubMed] [Google Scholar]
- Goelet P, Castellucci VF, Schacher S, Kandel ER. The long and the short of long-term memory–a molecular framework. Nature. 1986;322:419–422. doi: 10.1038/322419a0. [DOI] [PubMed] [Google Scholar]
- Gold PE. Acetylcholine modulation of neural systems involved in learning and memory. Neurobiology of Learning and Memory. 2003;80:194–210. doi: 10.1016/j.nlm.2003.07.003. [DOI] [PubMed] [Google Scholar]
- Gold PE. Glucose and age-related changes in memory. Neurobiology of Aging. 2005;26(Suppl 1):60–64. doi: 10.1016/j.neurobiolaging.2005.09.002. [DOI] [PubMed] [Google Scholar]
- Gold PE. The many faces of amnesia. Learning and Memory. 2006;13:506–514. doi: 10.1101/lm.277406. [DOI] [PubMed] [Google Scholar]
- Goodman RH, Flexner JB, Flexner LB. The effect of acetoxycycloheximide on rate of accumulation of cerebral catecholamines from circulating tyrosine as related to its effect on memory. Proceedings of the National Academy of Sciences of the United States of America. 1975;72:479–482. doi: 10.1073/pnas.72.2.479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Graves LA, Heller EA, Pack AI, Abel T. Sleep deprivation selectively impairs memory consolidation for contextual fear conditioning. Learning and Memory. 2003;10:168–176. doi: 10.1101/lm.48803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grecksch G, Matthies H. Two sensitive periods for the amnesic effect of anisomycin. Pharmacology, Biochemistry and Behavior. 1980;12:663–665. doi: 10.1016/0091-3057(80)90145-8. [DOI] [PubMed] [Google Scholar]
- Hagiwara M, Shimomura A, Yoshida K, Imaki J. Gene expression and CREB phosphorylation induced by cAMP and Ca2+ in neuronal cells. Advances in Pharmacology. 1996;36:277–285. doi: 10.1016/s1054-3589(08)60586-4. [DOI] [PubMed] [Google Scholar]
- Harada J, Sugimoto M. An inhibitor of p38 and JNK MAP kinases prevents activation of caspase and apoptosis of cultured cerebellar granule neurons. Japanese Journal of Pharmacology. 1999;79:369–378. doi: 10.1254/jjp.79.369. [DOI] [PubMed] [Google Scholar]
- Heasley LE, Storey B, Fanger GR, Butterfield L, Zamarripa J, Blumberg D, et al. GTPase-deficient G alpha 16 and G alpha q induce PC12 cell differentiation and persistent activation of cJun NH2-terminal kinases. Molecular and Cellular Biology. 1996;16:648–656. doi: 10.1128/mcb.16.2.648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hebb DO. The organization of behavior: A neuropsychological theory. New York: Wiley; 1949. [Google Scholar]
- Hellman K, Abel T. Fear conditioning increases NREM sleep. Behavioral Neuroscience. 2007;121:310–323. doi: 10.1037/0735-7044.121.2.310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hernandez PJ, Kelley AE. Long-term memory for instrumental responses does not undergo protein synthesis-dependent reconsolidation upon retrieval. Learning and Memory. 2004;11:748–754. doi: 10.1101/lm.84904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Holahan MR, Routtenberg A. Post-translational synaptic protein modification as substrate for long-lasting, remote memory: An initial test. Hippocampus. 2007;17:93–97. doi: 10.1002/hipo.20245. [DOI] [PubMed] [Google Scholar]
- Hong SS, Qian H, Zhao P, Bazzy-Asaad A, Xia Y. Anisomycin protects cortical neurons from prolonged hypoxia with differential regulation of p38 and ERK. Brain Research. 2007;1149:76–86. doi: 10.1016/j.brainres.2007.02.062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Horstmann S, Kahle PJ, Borasio GD. Inhibitors of p38 mitogen-activated protein kinase promote neuronal survival in vitro. Journal of Neuroscience Research. 1998;52:483–490. doi: 10.1002/(SICI)1097-4547(19980515)52:4<483::AID-JNR12>3.0.CO;2-4. [DOI] [PubMed] [Google Scholar]
- Huang EP. Synaptic plasticity: Going through phases with LTP. Current Biology. 1998;8:R350–R352. doi: 10.1016/s0960-9822(98)70219-2. [DOI] [PubMed] [Google Scholar]
- Huang YY, Martin KC, Kandel ER. Both protein kinase A and mitogen-activated protein kinase are required in the amygdala for the macromolecular synthesis-dependent late phase of long-term potentiation. Journal of Neuroscience. 2000;20:6317–6325. doi: 10.1523/JNEUROSCI.20-17-06317.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hunt E, Krivanek J. The effects of pentylenetetrazole and methylphenoxypropane on discrimination learning. Psychopharmacologia. 1966;9:1–16. doi: 10.1007/BF00427700. [DOI] [PubMed] [Google Scholar]
- Hyman SE, Malenka RC, Nestler EJ. Neural mechanisms of addiction: The role of reward-related learning and memory. Annual Review of Neuroscience. 2006 doi: 10.1146/annurev.neuro.29.051605.113009. [DOI] [PubMed] [Google Scholar]
- Igaz LM, Vianna MR, Medina JH, Izquierdo I. Two time periods of hippocampal mRNA synthesis are required for memory consolidation of fear-motivated learning. Journal of Neuroscience. 2002;22:6781–6789. doi: 10.1523/JNEUROSCI.22-15-06781.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Impey S, Obrietan K, Wong ST, Poser S, Yano S, Wayman G, et al. Cross talk between ERK and PKA is required for Ca2+ stimulation of CREB-dependent transcription and ERK nuclear translocation. Neuron. 1998;21:869–883. doi: 10.1016/s0896-6273(00)80602-9. [DOI] [PubMed] [Google Scholar]
- Iordanov MS, Pribnow D, Magun JL, Dinh TH, Pearson JA, Chen SL, et al. Ribotoxic stress response: Activation of the stress-activated protein kinase JNK1 by inhibitors of the peptidyl transferase reaction and by sequence-specific RNA damage to the alpha-sarcin/ricin loop in the 28S rRNA. Molecular and Cellular Biology. 1997;17:3373–3381. doi: 10.1128/mcb.17.6.3373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ip YT, Davis RJ. Signal transduction by the c-Jun N-terminal kinase (JNK)–from inflammation to development. Current Opinion in Cell Biology. 1998;10:205–219. doi: 10.1016/s0955-0674(98)80143-9. [DOI] [PubMed] [Google Scholar]
- Ji D, Wilson MA. Coordinated memory replay in the visual cortex and hippocampus during sleep. Nature Neuroscience. 2007;10:100–107. doi: 10.1038/nn1825. [DOI] [PubMed] [Google Scholar]
- Kaang BK, Kandel ER, Grant SG. Activation of cAMP-responsive genes by stimuli that produce long-term facilitation in Aplysia sensory neurons. Neuron. 1993;10:427–435. doi: 10.1016/0896-6273(93)90331-k. [DOI] [PubMed] [Google Scholar]
- Kalderon D, Rubin GM. Isolation and characterization of Drosophila cAMP-dependent protein kinase genes. Genes and Development. 1988;2:1539–1556. doi: 10.1101/gad.2.12a.1539. [DOI] [PubMed] [Google Scholar]
- Kalivas PW. How do we determine which drug-induced neuroplastic changes are important? Nature Neuroscience. 2005;8:1440–1441. doi: 10.1038/nn1105-1440. [DOI] [PubMed] [Google Scholar]
- Kandel ER. The molecular biology of memory storage: A dialogue between genes and synapses. Science. 2001;294:1030–1038. doi: 10.1126/science.1067020. [DOI] [PubMed] [Google Scholar]
- Kang H, Schuman EM. A requirement for local protein synthesis in neurotrophin-induced hippocampal synaptic plasticity. Science. 1996;273:1402–1406. doi: 10.1126/science.273.5280.1402. [DOI] [PubMed] [Google Scholar]
- Karpova A, Mikhaylova M, Thomas U, Knopfel T, Behnisch T. Involvement of protein synthesis and degradation in long-term potentiation of Schaffer collateral CA1 synapses. Journal of Neuroscience. 2006;26:4949–4955. doi: 10.1523/JNEUROSCI.4573-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kasten MR, Fan Y, Schulz PE. Activation of silent synapses with sustained but not decremental long-term potentiation. Neuroscience Letters. 2007;417:84–89. doi: 10.1016/j.neulet.2007.02.035. [DOI] [PubMed] [Google Scholar]
- Katz JJ, Kalstead WC. Protein organization and mental function. Comparative Psychology Monographs. 1950:1–38. [Google Scholar]
- Kelley AE. Memory and addiction; shared neural circuitry and molecular mechanisms. Neuron. 2004;44:161–179. doi: 10.1016/j.neuron.2004.09.016. [DOI] [PubMed] [Google Scholar]
- Kharlamov E, Cagnoli CM, Atabay C, Ikonomovic S, Grayson DR, Manev H. Opposite effect of protein synthesis inhibitors on potassium deficiency-induced apoptotic cell death in immature and mature neuronal cultures. Journal of Neurochemistry. 1995;65:1395–1398. doi: 10.1046/j.1471-4159.1995.65031395.x. [DOI] [PubMed] [Google Scholar]
- Kim JJ, DeCola JP, Landeira-Fernandez J, Fanselow MS. N-methyl-D-aspartate receptor antagonist APV blocks acquisition but not expression of fear conditioning. Behavioral Neuroscience. 1991;105:126–133. doi: 10.1037//0735-7044.105.1.126. [DOI] [PubMed] [Google Scholar]
- Kim EY, Mahmoud GS, Grover LM. REM sleep deprivation inhibits LTP in vivo in area CA1 of rat hippocampus. Neuroscience Letters. 2005;388:163–167. doi: 10.1016/j.neulet.2005.06.057. [DOI] [PubMed] [Google Scholar]
- Klann E, Dever TE. Biochemical mechanisms for translational regulation in synaptic plasticity. Nature Reviews of Neuroscience. 2004;5:931–942. doi: 10.1038/nrn1557. [DOI] [PubMed] [Google Scholar]
- Korzus E, Rosenfeld MG, Mayford M. CBP histone acetyltransferase activity is a critical component of memory consolidation. Neuron. 2004;42:961–972. doi: 10.1016/j.neuron.2004.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krug M, Lossner B, Ott T. Anisomycin blocks the late phase of long-term potentiation in the dentate gyrus of freely moving rats. Brain Research Bulletin. 1984;13:39–42. doi: 10.1016/0361-9230(84)90005-4. [DOI] [PubMed] [Google Scholar]
- Kudrimoti HS, Barnes CA, McNaughton BL. Reactivation of hippocampal cell assemblies: Effects of behavioral state, experience, and EEG dynamics. Journal of Neuroscience. 1999;19:4090–4101. doi: 10.1523/JNEUROSCI.19-10-04090.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kullmann DM, Lamsa KP. Long-term synaptic plasticity in hippocampal interneurons. Nat Rev Neurosci. 2007;8:687–699. doi: 10.1038/nrn2207. [DOI] [PubMed] [Google Scholar]
- Kurotani T, Higashi S, Inokawa H, Toyama K. Protein and RNA synthesis-dependent and -independent LTPs in developing rat visual cortex. Neuroreport. 1996;8:35–39. doi: 10.1097/00001756-199612200-00008. [DOI] [PubMed] [Google Scholar]
- Kyriakis JM, Avruch J. Sounding the alarm: Protein kinase cascades activated by stress and inflammation. Journal of Biological Chemistry. 1996;271:24313–24316. doi: 10.1074/jbc.271.40.24313. [DOI] [PubMed] [Google Scholar]
- Kyriakis JM, Woodgett JR, Avruch J. The stress-activated protein kinases. A novel ERK subfamily responsive to cellular stress and inflammatory cytokines. Annals of the New York Academy of Sciences. 1995;766:303–319. doi: 10.1111/j.1749-6632.1995.tb26683.x. [DOI] [PubMed] [Google Scholar]
- Lamprecht R, Hazvi S, Dudai Y. cAMP response element-binding protein in the amygdala is required for long- but not short-term conditioned taste aversion memory. Journal of Neuroscience. 1997;17:8443–8450. doi: 10.1523/JNEUROSCI.17-21-08443.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lashley KS. In search of the engram. Symposium of the Society for Experimental Biology. 1950;4:454–482. [Google Scholar]
- Lattal KM, Abel T. Different requirements for protein synthesis in acquisition and extinction of spatial preferences and context-evoked fear. Journal of Neuroscience. 2001;21:5773–5780. doi: 10.1523/JNEUROSCI.21-15-05773.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lechner HA, Squire LR, Byrne JH. 100 years of consolidation–remembering Muller and Pilzecker. Learning and Memory. 1999;6:77–87. [PubMed] [Google Scholar]
- Lee AK, Wilson MA. Memory of sequential experience in the hippocampus during slow wave sleep. Neuron. 2002;36:1183–1194. doi: 10.1016/s0896-6273(02)01096-6. [DOI] [PubMed] [Google Scholar]
- Le-Niculescu H, Bonfoco E, Kasuya Y, Claret FX, Green DR, Karin M. Withdrawal of survival factors results in activation of the JNK pathway in neuronal cells leading to Fas ligand induction and cell death. Molecular and Cellular Biology. 1999;19:751–763. doi: 10.1128/mcb.19.1.751. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leppa S, Eriksson M, Saffrich R, Ansorge W, Bohmann D. Complex functions of AP-1 transcription factors in differentiation and survival of PC12 cells. Molecular and Cellular Biology. 2001;21:4369–4378. doi: 10.1128/MCB.21.13.4369-4378.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Levenson JM, Sweatt JD. Epigenetic mechanisms: A common theme in vertebrate and invertebrate memory formation. Cellular and Molecular Life Sciences. 2006;63:1009–1016. doi: 10.1007/s00018-006-6026-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu J, Valencia-Sanchez MA, Hannon GJ, Parker R. MicroRNA-dependent localization of targeted mRNAs to mammalian P-bodies. Nat Cell Biol. 2005;7:719–723. doi: 10.1038/ncb1274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lopez-Mascaraque L, Price JL. Protein synthesis inhibitors delay transneuronal death in the piriform cortex of young adult rats. Neuroscience. 1997;79:463–475. doi: 10.1016/s0306-4522(96)00707-5. [DOI] [PubMed] [Google Scholar]
- Lynch GS, Dunwiddie T, Gribkoff V. Heterosynaptic depression: A postsynaptic correlate of long-term potentiation. Nature. 1977;266:737–739. doi: 10.1038/266737a0. [DOI] [PubMed] [Google Scholar]
- Mackiewicz M, Shockley KR, Romer MA, Galante RJ, Zimmerman JE, Naidoo N, et al. Macromolecule biosynthesis —a key function of sleep. Physiological Genomics. 2007 doi: 10.1152/physiolgenomics.00275.2006. [DOI] [PubMed] [Google Scholar]
- Malenka RC, Nicoll RA. Long-term potentiation–a decade of progress? Science. 1999;285:1870–1874. doi: 10.1126/science.285.5435.1870. [DOI] [PubMed] [Google Scholar]
- Maroney AC, Glicksman MA, Basma AN, Walton KM, Knight E, Jr, Murphy CA, et al. Motoneuron apoptosis is blocked by CEP-1347 (KT 7515), a novel inhibitor of the JNK signaling pathway. Journal of Neuroscience. 1998;18:104–111. doi: 10.1523/JNEUROSCI.18-01-00104.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martin KC. Local protein synthesis during axon guidance and synaptic plasticity. Current Opinion in Neurobiology. 2004;14:305–310. doi: 10.1016/j.conb.2004.05.009. [DOI] [PubMed] [Google Scholar]
- Martin KC, Casadio A, Zhu H, Yaping E, Rose JC, Chen M, et al. Synapse-specific, long-term facilitation of aplysia sensory to motor synapses: A function for local protein synthesis in memory storage. Cell. 1997;91:927–938. doi: 10.1016/s0092-8674(00)80484-5. [DOI] [PubMed] [Google Scholar]
- Martin SJ, Grimwood PD, Morris RG. Synaptic plasticity and memory: An evaluation of the hypothesis. Annual Review of Neuroscience. 2000;23:649–711. doi: 10.1146/annurev.neuro.23.1.649. [DOI] [PubMed] [Google Scholar]
- Martin KC, Kosik KS. Synaptic tagging—who’s it? Nat Rev Neurosci. 2002;3:813–820. doi: 10.1038/nrn942. [DOI] [PubMed] [Google Scholar]
- Martin KC, Zukin RS. RNA trafficking and local protein synthesis in dendrites: An overview. Journal of Neuroscience. 2006;26:7131–7134. doi: 10.1523/JNEUROSCI.1801-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matthies H. The biochemical basis of learning and memory. Life Sciences. 1974;15:2017–2031. doi: 10.1016/0024-3205(74)90019-8. [DOI] [PubMed] [Google Scholar]
- Matthies H. Neurobiological aspects of learning and memory. Annual Review of Psychology. 1989;40:381–404. doi: 10.1146/annurev.ps.40.020189.002121. [DOI] [PubMed] [Google Scholar]
- Matthies H, Frey U, Reymann K, Krug M, Jork R, Schroeder H. Different mechanisms and multiple stages of LTP. Advances in Experimental Medicine and Biology. 1990;268:359–368. doi: 10.1007/978-1-4684-5769-8_39. [DOI] [PubMed] [Google Scholar]
- Mayford M, Baranes D, Podsypanina K, Kandel ER. The 3′-untranslated region of CaMKII alpha is a cis-acting signal for the localization and translation of mRNA in dendrites. Proceedings of the National Academy of Sciences of the United States of America. 1996;93:13250–13255. doi: 10.1073/pnas.93.23.13250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McGaugh JL. Time-dependent processes in memory storage. Science. 1966;153:1351–1358. doi: 10.1126/science.153.3742.1351. [DOI] [PubMed] [Google Scholar]
- McGaugh JL. Hormonal influences on memory. Annual Review of Psychology. 1983;34:297–323. doi: 10.1146/annurev.ps.34.020183.001501. [DOI] [PubMed] [Google Scholar]
- McGaugh JL. The perseveration-consolidation hypothesis: Mueller and Pilzecker, 1900. Brain Research Bulletin. 1999;50:445–446. doi: 10.1016/s0361-9230(99)00126-4. [DOI] [PubMed] [Google Scholar]
- McGaugh JL. Memory–a century of consolidation. Science. 2000;287:248–251. doi: 10.1126/science.287.5451.248. [DOI] [PubMed] [Google Scholar]
- McGaugh JL, Petrinovich LF. Effects of drugs on learning and memory. International Review of Neurobiology. 1965;8:139–196. doi: 10.1016/s0074-7742(08)60757-6. [DOI] [PubMed] [Google Scholar]
- Miller RR, Matzel LD. Retrieval failure versus memory loss in experimental amnesia: Definitions and processes. Learning and Memory. 2006;13:491–497. doi: 10.1101/lm.241006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miller S, Yasuda M, Coats JK, Jones Y, Martone ME, Mayford M. Disruption of dendritic translation of CaMKII-alpha impairs stabilization of synaptic plasticity and memory consolidation. Neuron. 2002;36:507–519. doi: 10.1016/s0896-6273(02)00978-9. [DOI] [PubMed] [Google Scholar]
- Milner B. Disorders of learning and memory after temporal lobe lesions in man. Clinical Neurosurgery. 1972;19:421–446. doi: 10.1093/neurosurgery/19.cn_suppl_1.421. [DOI] [PubMed] [Google Scholar]
- Milner B, Squire LR, Kandel ER. Cognitive neuroscience and the study of memory. Neuron. 1998;20:445–468. doi: 10.1016/s0896-6273(00)80987-3. [DOI] [PubMed] [Google Scholar]
- Misanin JR, Miller RR, Lewis DJ. Retrograde amnesia produced by electroconvulsive shock after reactivation of a consolidated memory trace. Science. 1968;160:554–555. doi: 10.1126/science.160.3827.554. [DOI] [PubMed] [Google Scholar]
- Miyamoto E. Molecular mechanism of neuronal plasticity: Induction and maintenance of long-term potentiation in the hippocampus. Journal of Pharmacological Science. 2006;100:433–442. doi: 10.1254/jphs.cpj06007x. [DOI] [PubMed] [Google Scholar]
- Miyashiro K, Dichter M, Eberwine J. On the nature and differential distribution of mRNAs in hippocampal neurites: Implications for neuronal functioning. Proceedings of the National Academy of Sciences of the United States of America. 1994;91:10800–10804. doi: 10.1073/pnas.91.23.10800. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moccia R, Chen D, Lyles V, Kapuya E, E Y, Kalachikov S, et al. An unbiased cDNA library prepared from isolated Aplysia sensory neuron processes is enriched for cytoskeletal and translational mRNAs. Journal of Neuroscience. 2003;23:9409–9417. doi: 10.1523/JNEUROSCI.23-28-09409.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Molter C, Sato N, Yamaguchi Y. Reactivation of behavioral activity during sharp waves: A computational model for two stage hippocampal dynamics. Hippocampus. 2007;17:201–209. doi: 10.1002/hipo.20258. [DOI] [PubMed] [Google Scholar]
- Monné L. Functioning of the cytoplasm. Advanced Enzymology. 1948;8:1–69. [Google Scholar]
- Montarolo PG, Goelet P, Castellucci VF, Morgan J, Kandel ER, Schacher S. A critical period for macromolecular synthesis in long-term heterosynaptic facilitation in Aplysia. Science. 1986;234:1249–1254. doi: 10.1126/science.3775383. [DOI] [PubMed] [Google Scholar]
- Morgan JI, Curran T. Calcium as a modulator of the immediate-early gene cascade in neurons. Cell Calcium. 1988;9:303–311. doi: 10.1016/0143-4160(88)90011-5. [DOI] [PubMed] [Google Scholar]
- Morris RG. Synaptic plasticity and learning: Selective impairment of learning rats and blockade of long-term potentiation in vivo by the N-methyl-D-aspartate receptor antagonist AP5. Journal of Neuroscience. 1989;9:3040–3057. doi: 10.1523/JNEUROSCI.09-09-03040.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Müller GE, Pilzecker A. Experimentelle Beiträge zur Lehre vom Gedächtnis. Zeitschrift Fur Psychologie. 1900;1:1–288. [Google Scholar]
- Nader K. Memory traces unbound. Trends in Neurosciences. 2003;26:65–72. doi: 10.1016/S0166-2236(02)00042-5. [DOI] [PubMed] [Google Scholar]
- Nakajima S. Interference with relearning in the rat after hippocampal injection of actinomycin D. Journal of Comparative and Physiological Psychology. 1969;67:457–461. doi: 10.1037/h0027292. [DOI] [PubMed] [Google Scholar]
- Nakazawa K, McHugh TJ, Wilson MA, Tonegawa S. NMDA receptors, place cells and hippocampal spatial memory. Nat Rev Neurosci. 2004;5:361–372. doi: 10.1038/nrn1385. [DOI] [PubMed] [Google Scholar]
- Nathans D. Puromycin inhibition of protein synthesis: Incorporation of puromycin into peptide chains. Proceedings of the National Academy of Sciences of the United States of America. 1964;51:585–592. doi: 10.1073/pnas.51.4.585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nguyen PV, Abel T, Kandel ER. Requirement of a critical period of transcription for induction of a late phase of LTP. Science. 1994;265:1104–1107. doi: 10.1126/science.8066450. [DOI] [PubMed] [Google Scholar]
- Nguyen PV, Woo NH. Regulation of hippocampal synaptic plasticity by cyclic AMP-dependent protein kinases. Progress in Neurobiology. 2003;71:401–437. doi: 10.1016/j.pneurobio.2003.12.003. [DOI] [PubMed] [Google Scholar]
- Obata K, Yamanaka H, Kobayashi K, Dai Y, Mizushima T, Katsura H, et al. Role of mitogen-activated protein kinase activation in injured and intact primary afferent neurons for mechanical and heat hypersensitivity after spinal nerve ligation. Journal of Neuroscience. 2004;24:10211–10222. doi: 10.1523/JNEUROSCI.3388-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- O’Hara BF, Ding J, Bernat RL, Franken P. Genomic and proteomic approaches towards an understanding of sleep. CNS Neurol Disord Drug Targets. 2007;6:71–81. doi: 10.2174/187152707779940745. [DOI] [PubMed] [Google Scholar]
- Oike Y, Hata A, Mamiya T, Kaname T, Noda Y, Suzuki M, et al. Truncated CBP protein leads to classical Rubinstein-Taybi syndrome phenotypes in mice: Implications for a dominant-negative mechanism. Human Molecular Genetics. 1999;8:387–396. doi: 10.1093/hmg/8.3.387. [DOI] [PubMed] [Google Scholar]
- O’Keefe J, Nadel L. The hippocampus as a cognitive map. Oxford: Oxford University Press; 1978. [Google Scholar]
- Pare W. The effect of caffeine and seconal on a visual discrimination task. Journal of Comparative and Physiological Psychology. 1961;54:506–509. [PubMed] [Google Scholar]
- Pavlides C, Winson J. Influences of hippocampal place cell firing in the awake state on the activity of these cells during subsequent sleep episodes. Journal of Neuroscience. 1989;9:2907–2918. doi: 10.1523/JNEUROSCI.09-08-02907.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pestka S. Inhibitors of ribosome functions. Annual Review of Microbiology. 1971;25:487–562. doi: 10.1146/annurev.mi.25.100171.002415. [DOI] [PubMed] [Google Scholar]
- Quirk GJ. Learning not to fear, faster. Learning and Memory. 2004;11:125–126. doi: 10.1101/lm.75404. [DOI] [PubMed] [Google Scholar]
- Rainbow TC, Hoffman PL, Flexner LB. Studies of memory: A reevaluation in mice of the effects of inhibitors on the rate of synthesis of cerebral proteins as related to amnesia. Pharmacology, Biochemistry and Behavior. 1980;12:79–84. doi: 10.1016/0091-3057(80)90419-0. [DOI] [PubMed] [Google Scholar]
- Rao VR, Finkbeiner S. NMDA and AMPA receptors: Old channels, new tricks. Trends in Neurosciences. 2007;30:284–291. doi: 10.1016/j.tins.2007.03.012. [DOI] [PubMed] [Google Scholar]
- Rao A, Steward O. Evidence that protein constituents of postsynaptic membrane specializations are locally synthesized: Analysis of proteins synthesized within synaptosomes. Journal of Neuroscience. 1991;11:2881–2895. doi: 10.1523/JNEUROSCI.11-09-02881.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Raught B, Gingras AC, Sonenberg N. The target of rapamycin (TOR) proteins. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:7037–7044. doi: 10.1073/pnas.121145898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reijmers LG, Perkins BL, Matsuo N, Mayford M. Localization of a stable neural correlate of associative memory. Science. 2007;317:1230–1233. doi: 10.1126/science.1143839. [DOI] [PubMed] [Google Scholar]
- Reymann KG, Frey JU. The late maintenance of hippocampal LTP: Requirements, phases, ‘synaptic tagging’, ‘late-associativity’ and implications. Neuropharmacology. 2007;52:24–40. doi: 10.1016/j.neuropharm.2006.07.026. [DOI] [PubMed] [Google Scholar]
- Ribeiro S, Mello CV, Velho T, Gardner TJ, Jarvis ED, Pavlides C. Induction of hippocampal long-term potentiation during waking leads to increased extrahippocampal zif-268 expression during ensuing rapid-eye-movement sleep. Journal of Neuroscience. 2002;22:10914–10923. doi: 10.1523/JNEUROSCI.22-24-10914.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Richter JD, Lorenz LJ. Selective translation of mRNAs at synapses. Current Opinion in Neurobiology. 2002;12:300–304. doi: 10.1016/s0959-4388(02)00318-5. [DOI] [PubMed] [Google Scholar]
- Roberson ED, English JD, Adams JP, Selcher JC, Kondratick C, Sweatt JD. The mitogen-activated protein kinase cascade couples PKA and PKC to cAMP response element binding protein phosphorylation in area CA1 of hippocampus. Journal of Neuroscience. 1999;19:4337–4348. doi: 10.1523/JNEUROSCI.19-11-04337.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roberts RB, Flexner JB, Flexner LB. Some evidence for the involvement of adrenergic sites in the memory trace. Proceedings of the National Academy of Sciences of the United States of America. 1970;66:310–313. doi: 10.1073/pnas.66.2.310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rook MS, Lu M, Kosik KS. CaMKIIalpha 3′ untranslated region-directed mRNA translocation in living neurons: Visualization by GFP linkage. Journal of Neuroscience. 2000;20:6385–6393. doi: 10.1523/JNEUROSCI.20-17-06385.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roozendaal B, Carmi O, McGaugh JL. Adrenocortical suppression blocks the memory-enhancing effects of amphetamine and epinephrine. Proceedings of the National Academy of Sciences of the United States of America. 1996;93:1429–1433. doi: 10.1073/pnas.93.4.1429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roozendaal B, McGaugh JL. Amygdaloid nuclei lesions differentially affect glucocorticoid-induced memory enhancement in an inhibitory avoidance task. Neurobiology of Learning and Memory. 1996a;65:1–8. doi: 10.1006/nlme.1996.0001. [DOI] [PubMed] [Google Scholar]
- Roozendaal B, McGaugh JL. The memory-modulatory effects of glucocorticoids depend on an intact stria terminalis. Brain Research. 1996b;709:243–250. doi: 10.1016/0006-8993(95)01305-9. [DOI] [PubMed] [Google Scholar]
- Rosenzweig MR, Bennett EL, Colombo PJ, Lee DW, Serrano PA. Short-term, intermediate-term, and long-term memories. Behavioural Brain Research. 1993;57:193–198. doi: 10.1016/0166-4328(93)90135-d. [DOI] [PubMed] [Google Scholar]
- Routtenberg A, Rekart JL. Post-translational protein modification as the substrate for long-lasting memory. Trends in Neurosciences. 2005;28:12–19. doi: 10.1016/j.tins.2004.11.006. [DOI] [PubMed] [Google Scholar]
- Russell WR, Nathan PW. Traumatic amnesia. Brain. 1946;69:280–300. doi: 10.1093/brain/69.4.280. [DOI] [PubMed] [Google Scholar]
- Schaeffer HJ, Weber MJ. Mitogen-activated protein kinases: Specific messages from ubiquitous messengers. Molecular and Cellular Biology. 1999;19:2435–2444. doi: 10.1128/mcb.19.4.2435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schafe GE, Atkins CM, Swank MW, Bauer EP, Sweatt JD, LeDoux JE. Activation of ERK/MAP kinase in the amygdala is required for memory consolidation of pavlovian fear conditioning. Journal of Neuroscience. 2000;20:8177–8187. doi: 10.1523/JNEUROSCI.20-21-08177.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schneider AM, Sherman W. Amnesia: A function of the temporal relation of footshock to electroconvulsive shock. Science. 1968;159:219–221. doi: 10.1126/science.159.3811.219. [DOI] [PubMed] [Google Scholar]
- Schoenfeld RI, Seiden LS. Effect of alpha-methyltyrosine on operant behavior and brain catecholamine levels. Journal of Pharmacology and Experimental Therapeutics. 1969;167:319–327. [PubMed] [Google Scholar]
- Scholey AB, Rose SP, Zamani MR, Bock E, Schachner M. A role for the neural cell adhesion molecule in a late, consolidating phase of glycoprotein synthesis six hours following passive avoidance training of the young chick. Neuroscience. 1993;55:499–509. doi: 10.1016/0306-4522(93)90519-l. [DOI] [PubMed] [Google Scholar]
- Schwartzkroin PA, Wester K. Long-lasting facilitation of a synaptic potential following tetanization in the in vitro hippocampal slice. Brain Research. 1975;89:107–119. doi: 10.1016/0006-8993(75)90138-9. [DOI] [PubMed] [Google Scholar]
- Scoville WB, Milner B. Loss of recent memory after bilateral hippocampal lesions. Journal of Neurology, Neurosurgery and Psychiatry. 1957;20:11–21. doi: 10.1136/jnnp.20.1.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seger R, Krebs EG. The MAPK signaling cascade. FASEB Journal. 1995;9:726–735. [PubMed] [Google Scholar]
- Seiden LS, Peterson DD. Reversal of the reserpine-induced suppression of the conditioned avoidance response by L-DOPA: Correlation of behavioral and biochemical differences in two strains of mice. Journal of Pharmacology and Experimental Therapeutics. 1968;159:422–428. [PubMed] [Google Scholar]
- Serota RG, Roberts RB, Flexner LB. Acetoxycycloheximide-induced transient amnesia: Protective effects of adrenergic stimulants. Proceedings of the National Academy of Sciences of the United States of America. 1972;69:340–342. doi: 10.1073/pnas.69.2.340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sharp PE, McNaughton BL, Barnes CA. Enhancement of hippocampal field potentials in rats exposed to a novel, complex environment. Brain Research. 1985;339:361–365. doi: 10.1016/0006-8993(85)90105-2. [DOI] [PubMed] [Google Scholar]
- Shifrin VI, Anderson P. Trichothecene mycotoxins trigger a ribotoxic stress response that activates c-Jun N-terminal kinase and p38 mitogen-activated protein kinase and induces apoptosis. Journal of Biological Chemistry. 1999;274:13985–13992. doi: 10.1074/jbc.274.20.13985. [DOI] [PubMed] [Google Scholar]
- Silva AJ, Kogan JH, Frankland PW, Kida S. CREB and memory. Annual Review of Neuroscience. 1998;21:127–148. doi: 10.1146/annurev.neuro.21.1.127. [DOI] [PubMed] [Google Scholar]
- Skoulakis EM, Grammenoudi S. Dunces and da Vincis: The genetics of learning and memory in Drosophila. Cellular and Molecular Life Sciences. 2006;63:975–988. doi: 10.1007/s00018-006-6023-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sobell HM. Actinomycin and DNA transcription. Proceedings of the National Academy of Sciences of the United States of America. 1985;82:5328–5331. doi: 10.1073/pnas.82.16.5328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Soderling TR, Derkach VA. Postsynaptic protein phosphorylation and LTP. Trends in Neurosciences. 2000;23:75–80. doi: 10.1016/s0166-2236(99)01490-3. [DOI] [PubMed] [Google Scholar]
- Squire LR, Barondes SH. Actinomycin-D: Effects on memory at different times after training. Nature. 1970;225:649–650. doi: 10.1038/225649a0. [DOI] [PubMed] [Google Scholar]
- Squire LR, Barondes SH. Anisomycin, like other inhibitors of cerebral protein synthesis, impairs ‘long-term’ memory of a discrimination task. Brain Research. 1974;66:301–308. [Google Scholar]
- Squire LR, Barondes SH. Amnesic effect of cycloheximide not due to depletion of a constitutive brain protein with short half-life. Brain Research. 1976;103:183–189. doi: 10.1016/0006-8993(76)90703-4. [DOI] [PubMed] [Google Scholar]
- Squire LR, Bayley PJ. The neuroscience of remote memory. Current Opinion in Neurobiology. 2007;17:185–196. doi: 10.1016/j.conb.2007.02.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Squire LR, 2nd, Emanuel CA, Davis HP, Deutsch JA. Inhibitors of cerebral protein synthesis: Dissociation of aversive and amnesic effects. Behavioral Biology. 1975;14:335–341. doi: 10.1016/s0091-6773(75)90467-8. [DOI] [PubMed] [Google Scholar]
- Squire LR, Knowlton B, Musen G. The structure and organization of memory. Annual Review of Psychology. 1993;44:453–495. doi: 10.1146/annurev.ps.44.020193.002321. [DOI] [PubMed] [Google Scholar]
- Squire LR, Kuczenski R, Barondes SH. Tyrosine hydroxylase inhibition by cycloheximide and anisomycin is not responsible for their amnesic effect. Brain Research. 1974;82:241–248. doi: 10.1016/0006-8993(74)90601-5. [DOI] [PubMed] [Google Scholar]
- Staubli U, Faraday R, Lynch G. Pharmacological dissociation of memory: Anisomycin, a protein synthesis inhibitor, and leupeptin, a protease inhibitor, block different learning tasks. Behavioral and Neural Biology. 1985;43:287–297. doi: 10.1016/s0163-1047(85)91632-2. [DOI] [PubMed] [Google Scholar]
- Steward O, Levy WB. Preferential localization of polyribosomes under the base of dendritic spines in granule cells of the dentate gyrus. Journal of Neuroscience. 1982;2:284–291. doi: 10.1523/JNEUROSCI.02-03-00284.1982. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Steward O, Wallace CS, Lyford GL, Worley PF. Synaptic activation causes the mRNA for the IEG Arc to localize selectively near activated postsynaptic sites on dendrites. Neuron. 1998;21:741–751. doi: 10.1016/s0896-6273(00)80591-7. [DOI] [PubMed] [Google Scholar]
- Steward O, Worley P. Local synthesis of proteins at synaptic sites on dendrites: Role in synaptic plasticity and memory consolidation? Neurobiology of Learning and Memory. 2002;78:508–527. doi: 10.1006/nlme.2002.4102. [DOI] [PubMed] [Google Scholar]
- Stickgold R, Walker MP. Sleep-dependent memory consolidation and reconsolidation. Sleep Medicine. 2007;8:331–343. doi: 10.1016/j.sleep.2007.03.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stollhoff N, Menzel R, Eisenhardt D. Spontaneous recovery from extinction depends on the reconsolidation of the acquisition memory in an appetitive learning paradigm in the honeybee (Apis mellifera) Journal of Neuroscience. 2005;25:4485–4492. doi: 10.1523/JNEUROSCI.0117-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stork O, Welzl H. Memory formation and the regulation of gene expression. Cellular and Molecular Life Sciences. 1999;55:575–592. doi: 10.1007/s000180050316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stough S, Shobe JL, Carew TJ. Intermediate-term processes in memory formation. Current Opinion in Neurobiology. 2006;16:672–678. doi: 10.1016/j.conb.2006.10.009. [DOI] [PubMed] [Google Scholar]
- Stratton LO, Petrinovich L. Post-trial injections of an anti-cholinesterase drug and maze learning in two strains of rats. Psychopharmacologia. 1963;5:47–54. doi: 10.1007/BF00405574. [DOI] [PubMed] [Google Scholar]
- Sutton MA, Carew TJ. Parallel molecular pathways mediate expression of distinct forms of intermediate-term facilitation at tail sensory–motor synapses in Aplysia. Neuron. 2000;26:219–231. doi: 10.1016/s0896-6273(00)81152-6. [DOI] [PubMed] [Google Scholar]
- Sutton MA, Masters SE, Bagnall MW, Carew TJ. Molecular mechanisms underlying a unique intermediate phase of memory in Aplysia. Neuron. 2001;31:143–154. doi: 10.1016/s0896-6273(01)00342-7. [DOI] [PubMed] [Google Scholar]
- Sutton MA, Schuman EM. Dendritic protein synthesis, synaptic plasticity, and memory. Cell. 2006;127:49–58. doi: 10.1016/j.cell.2006.09.014. [DOI] [PubMed] [Google Scholar]
- Sutton MA, Wall NR, Aakalu GN, Schuman EM. Regulation of dendritic protein synthesis by miniature synaptic events. Science. 2004;304:1979–1983. doi: 10.1126/science.1096202. [DOI] [PubMed] [Google Scholar]
- Sweatt JD. The neuronal MAP kinase cascade: A biochemical signal integration system subserving synaptic plasticity and memory. Journal of Neurochemistry. 2001;76:1–10. doi: 10.1046/j.1471-4159.2001.00054.x. [DOI] [PubMed] [Google Scholar]
- Tang WJ, Gilman AG. Type-Specific regulation of adenylyl cyclase by G protein beta gamma subunits. Science. 1991;254:1500–1503. doi: 10.1126/science.1962211. [DOI] [PubMed] [Google Scholar]
- Tongiorgi E, Righi M, Cattaneo A. Activity-dependent dendritic targeting of BDNF and TrkB mRNAs in hippocampal neurons. Journal of Neuroscience. 1997;17:9492–9505. doi: 10.1523/JNEUROSCI.17-24-09492.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Torre ER, Steward O. Demonstration of local protein synthesis within dendrites using a new cell culture system that permits the isolation of living axons and dendrites from their cell bodies. Journal of Neuroscience. 1992;12:762–772. doi: 10.1523/JNEUROSCI.12-03-00762.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tronson NC, Taylor JR. Molecular mechanisms of memory reconsolidation. Nature Reviews of Neuroscience. 2007;8:262–275. doi: 10.1038/nrn2090. [DOI] [PubMed] [Google Scholar]
- Tully T, Preat T, Boynton SC, Del Vecchio M. Genetic dissection of consolidated memory in Drosophila. Cell. 1994;79:35–47. doi: 10.1016/0092-8674(94)90398-0. [DOI] [PubMed] [Google Scholar]
- Uphouse LL, MacInnes JW, Schlesinger K. Role of RNA and protein in memory storage: A review. Behavior Genetics. 1974;4:29–81. doi: 10.1007/BF01066705. [DOI] [PubMed] [Google Scholar]
- Vanhoutte P, Barnier JV, Guibert B, Pages C, Besson MJ, Hipskind RA, et al. Glutamate induces phosphorylation of Elk-1 and CREB, along with c-fos activation, via an extracellular signal-regulated kinase-dependent pathway in brain slices. Molecular and Cellular Biology. 1999a;19:136–146. doi: 10.1128/mcb.19.1.136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vanhoutte P, Barnier JV, Guibert B, Pages C, Besson MJ, Hipskind RA, et al. Glutamate induces phosphorylation of Elk-1 and CREB, along with c-fos activation, via an extracellular signal-regulated kinase-dependent pathway in brain slices. Molecular and Cellular Biology. 1999b;19:136–146. doi: 10.1128/mcb.19.1.136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vasquez D. Inhibitors of Protein Synthesis. New York: Springer-Verlag; 1979. [Google Scholar]
- Vecsey CG, Hawk JD, Lattal KM, Stein JM, Fabian SA, Attner MA, et al. Histone deacetylase inhibitors enhance memory and synaptic plasticity via CREB:CBP-dependent transcriptional activation. Journal of Neuroscience. 2007;27:6128–6140. doi: 10.1523/JNEUROSCI.0296-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Verkhratsky A, Kirchhoff F. NMDA Receptors in glia. Neuroscientist. 2007;13:28–37. doi: 10.1177/1073858406294270. [DOI] [PubMed] [Google Scholar]
- Walker MP. A refined model of sleep and the time course of memory formation. Behavioral and Brain Sciences. 2005;28:51–64. doi: 10.1017/s0140525x05000026. discussion 64–104. [DOI] [PubMed] [Google Scholar]
- Wang JQ, Fibuch EE, Mao L. Regulation of mitogenactivated protein kinases by glutamate receptors. Journal of Neurochemistry. 2007;100:1–11. doi: 10.1111/j.1471-4159.2006.04208.x. [DOI] [PubMed] [Google Scholar]
- Wang H, Hu Y, Tsien JZ. Molecular and systems mechanisms of memory consolidation and storage. Progress in Neurobiology. 2006;79:123–135. doi: 10.1016/j.pneurobio.2006.06.004. [DOI] [PubMed] [Google Scholar]
- Wang H, Storm DR. Calmodulin-regulated adenylyl cyclases: Cross-talk and plasticity in the central nervous system. Molecular Pharmacology. 2003;63:463–468. doi: 10.1124/mol.63.3.463. [DOI] [PubMed] [Google Scholar]
- Wang H, Tiedge H. Translational control at the synapse. Neuroscientist. 2004;10:456–466. doi: 10.1177/1073858404265866. [DOI] [PubMed] [Google Scholar]
- Warburton EC, Glover CP, Massey PV, Wan H, Johnson B, Bienemann A, et al. cAMP responsive element-binding protein phosphorylation is necessary for perirhinal long-term potentiation and recognition memory. Journal of Neuroscience. 2005;25:6296–6303. doi: 10.1523/JNEUROSCI.0506-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weiler IJ, Greenough WT. Metabotropic glutamate receptors trigger postsynaptic protein synthesis. Proceedings of the National Academy of Sciences of the United States of America. 1993;90:7168–7171. doi: 10.1073/pnas.90.15.7168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weiner N, Rabadjija M. The regulation of norepinephrine synthesis. Effect of puromycin on the accelerated synthesis of norepinephrine associated with nerve stimulation. Journal of Pharmacology and Experimental Therapeutics. 1968;164:103–114. [PubMed] [Google Scholar]
- Weisz DJ, Clark GA, Thompson RF. Increased responsivity of dentate granule cells during nictitating membrane response conditioning in rabbit. Behavioural Brain Research. 1984;12:145–154. doi: 10.1016/0166-4328(84)90037-8. [DOI] [PubMed] [Google Scholar]
- Wetzel W, Ott T, Matthies H. Is actinomycin D suitable for the investigation of memory processes? Pharmacology, Biochemistry and Behavior. 1976;4:515–519. doi: 10.1016/0091-3057(76)90190-8. [DOI] [PubMed] [Google Scholar]
- Whitlock JR, Heynen AJ, Shuler MG, Bear MF. Learning induces long-term potentiation in the hippocampus. Science. 2006;313:1093–1097. doi: 10.1126/science.1128134. [DOI] [PubMed] [Google Scholar]
- Wilson MA, McNaughton BL. Reactivation of hippocampal ensemble memories during sleep. Science. 1994;265:676–679. doi: 10.1126/science.8036517. [DOI] [PubMed] [Google Scholar]
- Wilson DA, Willner J, Kurz EM, Nadel L. Early handling increases hippocampal long-term potentiation in young rats. Behavioural Brain Research. 1986;21:223–227. doi: 10.1016/0166-4328(86)90240-8. [DOI] [PubMed] [Google Scholar]
- Wittstock S, Kaatz HH, Menzel R. Inhibition of brain protein synthesis by cycloheximide does not affect formation of long-term memory in honeybees after olfactory conditioning. Journal of Neuroscience. 1993;13:1379–1386. doi: 10.1523/JNEUROSCI.13-04-01379.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wittstock S, Menzel R. Color learning and memory in honey bees are not affected by protein synthesis inhibition. Behavioral and Neural Biology. 1994;62:224–229. doi: 10.1016/s0163-1047(05)80020-2. [DOI] [PubMed] [Google Scholar]
- Wood MA, Kaplan MP, Park A, Blanchard EJ, Oliveira AM, Lombardi TL, et al. Transgenic mice expressing a truncated form of CREB-binding protein (CBP) exhibit deficits in hippocampal synaptic plasticity and memory storage. Learning and Memory. 2005;12:111–119. doi: 10.1101/lm.86605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Worley PF, Cole AJ, Saffen DW, Baraban JM. Regulation of immediate early genes in brain: Role of NMDA receptor activation. Progress in Brain Research. 1990;86:277–285. doi: 10.1016/s0079-6123(08)63184-2. [DOI] [PubMed] [Google Scholar]
- Wu H, Zhou Y, Xiong ZQ. Transducer of regulated CREB and late phase long-term synaptic potentiation. Febs J. 2007;274:3218–3223. doi: 10.1111/j.1742-4658.2007.05891.x. [DOI] [PubMed] [Google Scholar]
- Xia Z, Dickens M, Raingeaud J, Davis RJ, Greenberg ME. Opposing effects of ERK and JNK-p38 MAP kinases on apoptosis. Science. 1995;270:1326–1331. doi: 10.1126/science.270.5240.1326. [DOI] [PubMed] [Google Scholar]
- Yin JC, Tully T. CREB and the formation of long-term memory. Current Opinion in Neurobiology. 1996;6:264–268. doi: 10.1016/s0959-4388(96)80082-1. [DOI] [PubMed] [Google Scholar]
- Yin JC, Wallach JS, Del Vecchio M, Wilder EL, Zhou H, Quinn WG, et al. Induction of a dominant negative CREB transgene specifically blocks long-term memory in Drosophila. Cell. 1994;79:49–58. doi: 10.1016/0092-8674(94)90399-9. [DOI] [PubMed] [Google Scholar]
- Zheng S, Zuo Z. Isoflurane preconditioning induces neuroprotection against ischemia via activation of P38 mitogen-activated protein kinases. Molecular Pharmacology. 2004;65:1172–1180. doi: 10.1124/mol.65.5.1172. [DOI] [PubMed] [Google Scholar]
- Zhong J, Zhang T, Bloch LM. Dendritic mRNAs encode diversified functionalities in hippocampal pyramidal neurons. BMC Neuroscience. 2006;7:17. doi: 10.1186/1471-2202-7-17. [DOI] [PMC free article] [PubMed] [Google Scholar]