CrebB-17A


DEVELOPMENTAL BIOLOGY

Larval and Adult

There is a complex pattern of transcripts apparent in larval stages, and in the heads and bodies of adult flies, with at least 12 different size transcripts apparent. The adult head contains at least six transcripts (Yin 1994).

Adult Fmr1 mutant flies display arrhythmic circadian activity and have erratic patterns of locomotor activity, whereas overexpression of Fmr1 leads to a lengthened period. Fmr1 mutant males also display reduced courtship activity which appears to result from their inability to maintain courtship interest. Molecular analysis fails to reveal any defects in the expression of clock components; however, the CREB output is affected. Morphological analysis of neurons required for normal circadian behavior reveals subtle abnormalities, suggesting that defects in axonal pathfinding or synapse formation may cause the observed behavioral defects (Dockendorff, 2002).

One known clock-controlled gene in Drosophila is the cAMP response element binding protein (CREB). To determine if the circadian oscillation of this protein is affected in the Fmr1 mutant flies, Fmr1 mutant flies carrying the CRE-luciferase (CRE-luc) reporter gene were examined in a luminometer continuously in constant daylight (DD) for up to 4 days. Although cycling of the CRE-luc reporter is detected in the Fmr1 mutant background, the amplitude of the oscillations is clearly reduced compared to the oscillations in the control background. This result indicates that dfmr1 affects a known molecular output of the clock. Normal cycling of PDF levels was seen in the termini of the small lateral neurons in the Fmr1 mutant brains. Thus this output of the clock is not affected at the normal site of its release, providing further evidence for normal central clock functioning in the Fmr1 mutant flies (Dockendorff, 2002).

Ecdysone signaling regulates the formation of long-term courtship memory in adult Drosophila melanogaster

Improved survival is likely linked to the ability to generate stable memories of significant experiences. Considerable evidence in humans and mammalian model animals shows that steroid hormones, which are released in response to emotionally arousing experiences, have an important role in the consolidation of memories of such events. In insects, ecdysone is the major steroid hormone, and it is well characterized with respect to its essential role in coordinating developmental transitions such as larval molting and metamorphosis. However, the functions of ecdysone in adult physiology remain largely elusive. This study shows that 20-hydroxyecdysone (20E), the active metabolite of ecdysone that is induced by environmental stimuli in adult Drosophila, has an important role in the formation of long-term memory (LTM). In male flies, the levels of 20E were found to be significantly increased after courtship conditioning, and exogenous administration of 20E either enhanced or suppressed courtship LTM, depending on the timing of its administration. Mutants in which ecdysone signaling is reduced are defective in LTM, and an elevation of 20E levels is associated with activation of the cAMP response element binding protein (CREB), an essential regulator of LTM formation. These results demonstrate that the molting steroid hormone ecdysone in adult Drosophila is critical to the evolutionarily conserved strategy that is used for the formation of stable memories. It is proposed that ecdysone is able to consolidate memories possibly by recapturing molecular and cellular processes that are used for normal neural development (Ishimoto, 2009).

The objective of this study was to investigate whether the steroid molting hormone 20E regulates LTM formation in adult Drosophila. This study shows the following; (1) training for courtship-memory leads to an elevation of 20E levels in adult flies; (2) administering exogenous 20E has either a positive or negative effect on courtship LTM, depending on the context; (3) disrupting either ecdysone synthesis or function of the nuclear EcR results in defective LTM; (4) functional ecdysone signaling in adult neurons during the training period is required for LTM, and (5) 20E induces CREB-mediated transcriptional activation. Together, these results indicate that the steroid molting hormone 20E has a novel, nondevelopmental role in the formation of long-lasting memory in adult insects (Ishimoto, 2009),

The temporal profile of 20E titers during embryonic, larval, and pupal stages is essentially controlled by the genetically determined developmental program. As previously shown, environmental stimuli, such as high temperature and nutritional shortage, induce up-regulation of 20E levels in adult flies. This study has demonstrated that 20E levels are increased in male flies after they are paired with a mated female for 7 h, conditions under which a robust courtship LTM is generated. Ecdysone signaling activated by these environmental stimuli or social interactions may trigger specific molecular and cellular responses in adults, and lead to long-lasting changes in physiology and behavior (Ishimoto, 2009),

In flies, steroid hormone synthesis is known to occur primarily in 2 organs, the larval prothoracic gland and the adult female ovary . Ecdysteroids are present in adult males as well as females. It remains to be determined where ecdysteroids are produced other than in the female ovary, and how their synthesis is regulated in adults. The last 4 sequential hydroxylations of their synthesis, which convert steroid precursors into 20E, are catalyzed by 4 cytochrome P450 enzymes encoded by phantom, disembodied, shadow, and shade, known collectively as the Halloween genes. The temporal changes in ecdysteroid levels during development are mainly attributed to transcriptional regulation of these genes. To understand the regulatory mechanisms for production of ecdysteroids in adult flies, it is important to examine where these enzymes are expressed, and how their expression and activity are regulated. Recent studies show that feeding the dopamine precursor L-DOPA to young Drosophila virilis females increases the dopamine (DA) content in the body, and subsequently results in a substantial increase in 20E levels. Given that dopamine has been implicated in negatively reinforced memory, it is possible that this neurotransmitter acts as a mediator between environmental stimuli and an elevation of 20E level (Ishimoto, 2009),

Using a temperature-sensitive EcR allele and an RNAi that targets EcR, it was shown that courtship LTM is impaired by conditional suppression of EcR function during the training period. Also, LTM was restored in the EcR temperature-sensitive mutants as long as they were maintained at the permissive temperature during the training period. These experiments demonstrate that ecdysone signaling through nuclear EcRs has an important role in the physiological processes that are necessary for the formation of LTM. How does ecdysone contribute to the formation of LTM? One possibility is that fully functional ecdysone signaling is required for effective sensory processing, and that the adverse effect of a 50% reduction in EcR expression on the learning process is due to severe sensory dysfunction. However, this possibility is not likely, because the courtship behavior of male flies with reduced EcR function was fond to be qualitatively and quantitatively comparable with that of control males. Also, EcR/+ males exhibited a short-lasting courtship memory after 1-h training, which suggests that their sensory acuity and ability to acquire courtship memory are rather normal. Thus, it is proposed that ecdysone signaling operates in the CNS, and contributes to consolidation of the memories into a long-lasting form. The MB is considered to be the center of olfactory memory. The EcR RNAi experiments suggest that the MB is one of the brain structures required for the influence of ecdysone on the formation of courtship LTM. Also, the study using the CRE-luc reporter indicates that CREB, a key regulator of long-lasting modifications of the nervous system, is involved in ecdysone-dependent LTM formation (Ishimoto, 2009),

Given that genetically programmed ecdysone signaling is known to control neuronal remodeling during development, it is interesting to speculate that certain experiences may recapture the ecdysone-mediated developmental processes in the adult brain and lead to structural and functional modifications to the nervous system that facilitate the formation of stable, LTM. The ability of ecdysone to remodel the nervous system is known not to be limited to developmental stages. For example, in the adult house cricket (Acheta domesticus) brain, ecdysone has been shown to inhibit proliferation of neuroblasts in the MBs and to trigger their differentiation into interneurons. Although there is no evidence of continued neurogenesis in the adult Drosophila brain, it is possible that ecdysone signaling induces significant changes in properties of existing neurons, resulting in structural and functional remodeling of neuronal circuits. A recent study has shown that the canonical ecdysteroid transcriptional cascade in the MB neurons of the adult worker honey bee (Apis mellifera) is initiated in response to activated ecdysone signaling, further suggesting the involvement of ecdysteroids in remodeling the adult nervous system (Ishimoto, 2009),

These findings in Drosophila indicate that regulation of memory by environmentally induced steroids could be ancient in origin, and widespread in species that have an ability to learn and remember. Thus, the molecular components and signaling pathways responsible for steroid-mediated memory regulation are likely to be shared, at least in part, by evolutionarily diverse animal species. This study has focused on the role of EcRs, nuclear hormone receptors that function through transcriptional regulation of their target genes, in the formation of LTM. Recently, a novel Drosophila G protein-coupled receptor (DopEcR) was found to be activated by ecdysteroids. Thus, it is also interesting to examine the possible involvement of rapid, nongenomic actions of ecdysone in regulation of memory. Considering the relatively simple nervous system of flies, the extensive knowledge of the genetics of this organism, and the highly developed experimental tools available for its study, Drosophila should be an ideal model system to elucidate the molecular, cellular, and neural-circuit bases of memory regulation by steroid hormones (Ishimoto, 2009),

Fasting launches CRTC to facilitate long-term memory formation in Drosophila

Canonical aversive long-term memory (LTM) formation in Drosophila requires multiple spaced trainings, whereas appetitive LTM can be formed after a single training. Appetitive LTM requires fasting prior to training, which increases motivation for food intake. However, this study found that fasting facilitates LTM formation in general; aversive LTM formation also occurred after single-cycle training when mild fasting was applied before training. Both fasting-dependent LTM (fLTM) and spaced training-dependent LTM (spLTM) requires protein synthesis and cyclic adenosine monophosphate response element-binding protein (CREB) activity. However, spLTM requires CREB activity in two neural populations--mushroom body and dorsal-anterior-lateral (DAL) neurons--whereas fLTM required CREB activity only in mushroom body neurons. fLTM uses the CREB coactivator CREB-regulated transcription coactivator (CRTC, whereas spLTM uses the coactivator CBP. Thus, flies use distinct LTM machinery depending on their hunger state (Hirano, 2013).

In Drosophila, canonical aversive long-term memory (LTM), which is dependent on de novo gene expression and protein synthesis, is generated after multiple rounds of spaced training. In contrast, appetitive LTM can be formed by single-cycle training. Because both aversive and appetitive LTM require protein synthesis and activation of CREB, it is likely that both types of LTM are formed by similar mechanisms. Appetitive and aversive LTM are known to differ (i.e., octopamine is specifically involved in appetitive but not aversive memory formation). However, it remains unclear why single-cycle training is sufficient for appetitive but not aversive LTM formation. Appetitive LTM cannot form unless fasting precedes training. Although fasting increases motivation for food intake (a requirement for appetitive memory) it was suspected that fasting may activate a second, motivation-independent, memory mechanism that facilitates LTM formation after single-cycle training (Hirano, 2013).

Flies were deprived of food for various periods of time and then subjected to aversive single-cycle training. Fasting prior to training significantly enhanced 1-day memory, with a peak at 16 hours of fasting and a return to nonfasting levels at 20 to 24 hours of fasting. In contrast, 16 hours of fasting did not increase short-term memory (STM, measured 1 hour after training). In this protocol, flies were returned to food vials after training, raising a possibility that the perception of food as a reward after training may enhance the previous aversive memory. This possibility was tested by inserting refeeding periods between food deprivation and training. Although fasting followed by a 4-hour refeeding period failed to induce appetitive LTM, it significantly enhanced aversive 1-day memory; this finding suggests that enhancement of aversive memory occurs through a mechanism unrelated to increased motivation or perception of food as a reward. A 6-hour refeeding period was sufficient to prevent aversive memory enhancement. Continuous food deprivation after training suppressed aversive memory enhancement, which indicates that both fasting before training and feeding after training are required to enhance aversive memory (Hirano, 2013).

Administration of the protein synthesis inhibitor cycloheximide (CHX) abolished 1-day memory enhancement but had no effect on 1-hour memory, supporting the idea that memory enhancement consists of an increase of LTM. Memory remaining after CHX treatment is likely to be protein synthesis-independent, anesthesia-resistant memory (ARM). Fasting for 16 hours neither enhanced protein synthesis-independent memory nor canonical aversive LTM generated by spaced training (spLTM). Furthermore, fasting-dependent memory decayed within 4 days, and food deprivation did not enhance 4-day spLTM, indicating that fasting-dependent memory is physiologically different from spLTM (Hirano, 2013).

Fasting-dependent memory was blocked by acute, dose-dependent, expression of CREB2-b, a repressor isoform of CREB, in the mushroom bodies (MBs). Expression of the repressor from two copies of UAS-CREB2-b under control of the MB247-Switch (MBsw) GAL4 driver, which induces UAS transgene expression upon RU486 feeding, significantly suppressed fasting-dependent memory upon RU486 feeding, whereas expression from one copy of UAS-CREB2-b did not. Defects in LTM formation are highly correlated with CREB2-b amounts. Significantly higher MBsw-dependent expression of CREB proteins was found in flies carrying two copies of UAS-CREB2-b relative to flies carrying one copy. MBsw-dependent CREB2-b expression did not affect STM in either fed or food-deprived conditions. Because the aversive memory enhanced by fasting is mediated by protein synthesis and CREB, this memory is referred to as fasting-dependent LTM (fLTM). Similar to the results in aversive fLTM, MBsw-dependent CREB2-b expression also decreased appetitive LTM but not appetitive STM (Hirano, 2013).

A recent study (Chen, 2012) concluded that CREB activity in MB neurons is not required for spLTM. In that study, CREB2-b was expressed using the OK107 MB driver and GAL80ts was used to restrict CREB2-b expression to 30°C. However, this study found that the GAL80ts construct still inhibited expression of CREB considerably at 30°C. When higher amounts of CREB2-b were acutely expressed in MBs using MBsw, a significant decrease was observed in 1-day spLTM, indicating that CREB activity in the MBs is likely to be required for spLTM (Hirano, 2013). Consistent with the results of Chen (2012) expression of CREB2-b in two dorsal-anterior-lateral (DAL) neurons impaired aversive spLTM. In contrast, expression of CREB2-b in DAL neurons did not affect aversive fLTM. Moreover, appetitive LTM was also not affected by expression of CREB2-b in DAL neurons. MBsw did not express GAL4 in DAL neurons (Hirano, 2013).

CREB requires coactivators, including CBP (CREB-binding protein), to activate transcription needed for LTM formation. Acute expression of an inverted repeat of CBP (CBP-IR) in MBs significantly impaired spLTM without affecting either STM or 1-day memory after multiple massed trainings, which do not lead to LTM formation. However, neither aversive fLTM nor appetitive LTM was impaired by CBP-IR expression, indicating that an alternative coactivator may be required for fasting-dependent memory (Hirano, 2013).

Recent studies demonstrate the involvement of a cAMP-regulated transcriptional coactivator (CRTC) in hippocampal plasticity (Kovacs, 2007; Zhou, 2006). In metabolic tissues, phosphorylated CRTC is sequestered in the cytoplasm while dephosphorylated CRTC translocates to the nucleus to promote CREB-dependent gene expression. Fasting causes CRTC dephosphorylation and activation. In line with this, significant accumulation of hemagglutinin (HA)-tagged CRTC (CRTC-HA) was found within MB nuclei after 16 hours of food deprivation. Subcellular fractionation indicated that food deprivation causes CRTC-HA nuclear translocation without affecting total CRTC-HA amounts (Hirano, 2013).

To examine the role of CRTC in fLTM and spLTM, a CRTC inverted repeat (CRTC-IR) was acutely expressed using MBsw, and suppression of aversive fLTM was observed but no effect was seen on STM. CHX treatment did not further decrease 1-day aversive memory, and CRTC-IR expression from a second MB driver, OK107, also impaired fLTM formation. CRTC-IR expression from MBsw also impaired appetitive LTM without affecting appetitive STM. In contrast, CRTC-IR expression from MBsw did not impair spLTM . CRTC-IR expression in DAL neurons had no effect on either aversive fLTM or appetitive LTM. Consistent with these results showing lack of fLTM after 24-hour fasting, 1-day aversive memory after 24-hour fasting did not decrease upon CRTC-IR expression in MBs (Hirano, 2013).

To examine the effects of spaced training on fLTM and the effects of fasting on spLTM, fed or fasted flies expressing either CBP-IR or CRTC-IR were space-trained. When CBP-IR was expressed to impair spLTM, 1-day memory after spaced training was impaired in fed conditions but not in fasting conditions, which suggested that spaced training protocols do not block fLTM. When CRTC-IR was expressed to impair fLTM formation, 1-day memory after spaced training was not affected by fasting, which suggested that mild fasting does not impair spLTM formation (Hirano, 2013).

Is activation of CRTC sufficient to generate fLTM in the absence of fasting? HA-tagged constitutively active CRTC (CRTC-SA-HA) was expressed from MBsw, and its nuclear accumulation was observed in the absence of fasting. Acute expression of CRTC-SA-HA from MBsw increased 1-day aversive memory after single-cycle training in fed flies, and this increase was not further enhanced by fasting. In contrast, expression of control CRTC-HA did not alter the fasting requirement for memory enhancement. CRTC-SA-HA expression did not affect feeding itself, which suggested that the memory enhancement is not due to impaired feeding. Taken together, CRTC activity in MBs is necessary and sufficient to form fLTM. Similar to the effects of fasting, CRTC-SA-HA expression did not affect STM or 4-day spLTM (Hirano, 2013).

In mammalian metabolic tissues, CRTC is phosphorylated by insulin signaling, which is suppressed by fasting (see Wang, 2008). CRTC phosphorylation is also regulated by insulin signaling in flies (Wang, 2008). To determine whether reduced insulin signaling activates CRTC and promotes fLTM formation, heterozygous mutants for chico, which encodes an adaptor protein required for insulin signaling, were tested. Although chico1 null mutants are semilethal and defective for olfactory learning, heterozygous chico1/+ mutants are viable and display normal learning (Hirano, 2013).

CRTC accumulated in MB nuclei in chico1/+ mutants in the absence of food deprivation. Under conditions where flies were fed, chico1/+ flies had significantly greater 1-day memory after single-cycle training relative to control flies, whereas 1-hour memory was unaffected. Enhanced 1-day memory in chico1/+ flies was not further enhanced by fasting. Because the chico1/+ mutation does not affect feeding itself, the memory enhancement would not seem to be attributable to impaired feeding. The increased 1-day memory in chico1/+ mutants was suppressed by CHX treatment and CRTC-IR expression using MBsw, which suggests that reduced insulin signaling mimics fLTM through activation of CRTC in MBs (Hirano, 2013).

Single-cycle training after mild fasting generates both appetitive and aversive LTM, and CRTC in the MBs plays a key role in both types of LTM. A CRTC-dependent LTM pathway is unlikely to be involved in increasing motivation required to form appetitive memory, because CRTC knockdown did not affect appetitive STM and because CRTC-SA expression was not sufficient to form appetitive LTM without prior fasting. Although mild 16-hour fasting induced aversive fLTM, longer 24-hour fasting impaired aversive fLTM but not appetitive LTM. Thus, although aversive and appetitive fLTM share mechanistic similarities, they may be regulated by different inputs controlling motivation and fasting time courses. Because nuclear translocation of CRTC was sustained even after 24 hours of food deprivation, prolonged fasting may suppress a CRTC-independent step in aversive fLTM formation. spLTM was not affected by 24-hour fasting prior to training, which suggests that the unknown inhibitory effect of 24-hour fasting does not occur after spaced training. Continuous food deprivation after training suppressed aversive fLTM. In another study, Placais (2013) reports that continuous food-deprivation after spaced training suppresses spLTM as well (Hirano, 2013).

Suppression of aversive LTM by prolonged fasting may ensure that starving flies pursue available food, with less concern for safety. Although the biological importance of aversive fLTM in natural environments is currently unclear, the current results indicate that different physiological states may induce different types of LTM in flies (Hirano, 2013).

Spatio-temporal in vivo recording of dCREB2 dynamics in Drosophila long-term memory processing

CREB (cAMP response element-binding protein) is an evolutionarily conserved transcription factor, playing key roles in synaptic plasticity, intrinsic excitability and long-term memory (LTM) formation. The Drosophila homologue of mammalian CREB, dCREB2, is also important for LTM. However, the spatio-temporal nature of dCREB2 activity during memory consolidation is poorly understood. Using an in vivo reporter system, this study examined dCREB2 activity continuously in specific brain regions during LTM processing. Two brain regions that have been shown to be important for Drosophila LTM are the ellipsoid body (EB) and the mushroom body (MB). dCREB2 reporter activity is persistently elevated in EB R2/R4m neurons, but not neighboring R3/R4d neurons, following LTM-inducing training. In multiple subsets of MB neurons, dCREB2 reporter activity is suppressed immediately following LTM-specific training, and elevated during late windows. In addition, heterogeneous responses were observed across different subsets of neurons in MB αβ lobe during LTM processing. All of these changes suggest that dCREB2 functions in both the EB and MB for LTM formation, and that this activity contributes to the process of systems consolidation (Zhang, 2014).

Effects of Mutation

In Drosophila, rest shares features with mammalian sleep, including prolonged immobility, decreased sensory responsiveness and a homeostatic rebound after deprivation. To understand the molecular regulation of sleep-like rest, the involvement of a candidate gene, cAMP response-element binding protein (CREB), was investigated. The duration of rest is inversely related to cAMP signaling and CREB activity. Acutely blocking CREB activity in transgenic flies does not affect the clock, but increases rest rebound. CREB mutants also have a prolonged and increased homeostatic rebound. In wild types, in vivo CREB activity increases after rest deprivation and remains elevated for a 72-hour recovery period. These data indicate that cAMP signaling has a non-circadian role in waking and rest homeostasis in Drosophila (Hendricks, 2001).

The daily rest of flies carrying mutations and/or transgenes that alter cAMP signaling was examined at several points in the pathway. dunce flies have a mutation in the phosphodiesterase enzyme and therefore have increased cAMP. The null mutant (dncML) rests significantly less than the background yw strain. Similarly, increasing PKA activity in flies with a heat-shock-inducible transgene of the catalytic subunit of PKA significantly decreases daily rest durations compared to pre-heat-shock rest levels. Decreased adenylyl cyclase enzyme activity and thus decreased cAMP characterize rutabaga (rut) mutants, which rest more than the Canton S background strain. Similarly, S162 flies that carry a mutation that abolishes dCREB2 activity rest more than their comparison group (siblings without the mutation). The mutation is a stop codon just upstream of the basic leucine-zipper motif of the dCREB2 gene (Hendricks, 2001).

Lines of flies with the heat shock-inducible activating (HS-dCREB2a) and blocking (HS-dCREB2b) dCREB2 transgenes were also examined. dCREB2 is a major target of PKA in Drosophila, and these transgenes have effects on long-term memory consolidation in Drosophila. Even without heat shock, baseline rest is increased in flies carrying the HS-dCREB2b transgene, whereas the flies with the inducible activator rest slightly but significantly less, suggesting a leaky heat shock promoter. When the locomotor activity was measured on the same days in all of these lines, three measures of daily activity were not significantly correlated with rest levels, providing evidence that rest is regulated independently of locomotor activity, and that the increase in rest with decreasing cAMP signaling is not due to general debility or sluggishness (Hendricks, 2001).

Because rest is inversely related to the level of cAMP signaling or dCREB2 activity, it seemed that the normal dCREB2 peak might be important for the animal to maintain normal waking. That is, nighttime dCREB2 might have a function for subsequent waking, consistent with the idea that dCREB2 might mediate a restorative function of rest, permitting or fostering sustained waking. Wild-type flies respond to six hours of rest deprivation at night by exhibiting a rest rebound (an increase in rest duration) for the morning six hours of each day of a three-day recovery period. This rebound is related to the duration of rest deprivation, and is not seen when the flies are subjected to the same stimulation during their usual active period. If the nocturnal peak in dCREB2 were necessary for recovery from rest deprivation, blocking the normal dCREB2 activity peak would be expected to impair the ability to recover after rest deprivation, as nighttime dCREB2-dependent gene expression would be abolished. In contrast, overexpressing the activator just before the usual peak might have minimal effects if the normal CREB-mediated transcription is already sufficient for normal function (Hendricks, 2001).

The response to rest deprivation was studied in the wild-type isogenic background strain, flies with the blocker transgene (HS-dCREB2b), and flies with the activator transgene (HS-dCREB2a). Wild-type and transgenic dCREB2 flies were deprived of six hours of rest, with or without heat shock. For each genotype, rest-deprived flies were compared to flies that were allowed to rest undisturbed, and to controls that were subjected to handling but were not rest-deprived. The mixed-model analysis of variance takes into account both between- and within-animal factors. The within-animal factor in this case is the pattern of each individual's rest throughout the study, and between-animal factors are the effects of experimental group, genotype and heat shock. A significant interaction (p < 0.0001) existed in the effect of heat shock and experimental group (resting, handled controls or deprived) genotype. That is, the effect of heat shock on rest duration depended on both the experimental condition and the genotype. By using a series of specific post hoc comparisons using only the data from after heat shock, it was found that inducing the blocker isoform of dCREB2 specifically increases the rest of these flies during recovery from deprivation throughout the entire period after deprivation. In contrast, inducing the activator (dCREB2a) does not significantly alter the rest rebound over the three-day period after deprivation. Heat shock alone does not increase rest in the undisturbed or handled control dCREB2b flies. Thus, rest rebound is enhanced only when dCREB2b induction is combined with rest deprivation. The increased rest during recovery from rest deprivation is detectable in individual dCREB2b flies as well as in the populations. The mean daily rest rebound in HS-dCREB2b flies is increased after deprivation on all three successive days, although the degree of rebound falls over time (from 1.96 hours above baseline on the first day to only 0.67 hour on the third day after deprivation). The ability to move is not differentially changed by heat shock and deprivation in dCREB2b flies compared to wild-type flies, as measured by changes in peak activity during the period after deprivation (Hendricks, 2001).

Because CREB is involved in responses to stress in several systems, blocking CREB activity may somehow alter the flies' response to six hours of stimulation, independent of any rest-related function. The response of dCREB2b flies was studied to the same combination of six hours stimulation and heat shock, applied during the usual daytime active period (heat shock at circadian time [CT] 0, stimulation from CT 6 to 12). The response (change in rest compared to baseline and to handled controls) of dCREB2b and background flies was statistically the same (Hendricks, 2001).

The findings that blocking dCREB2 increases rest rebound, and that rest rebound is associated with an increase in CRE-dependent gene expression, implicate CREB activity in a restorative function of rest. It is hypothesized that CREB activation during rest optimizes waking CNS function. In this context, it is interesting that cAMP signaling, PKA and CREB activity have a conserved role in learning and memory. One putative restorative function of sleep, optimizing neural plasticity, could be evolutionarily ancient. To directly test whether CREB also has a conserved role in states of arousal, CREB mutant mice have been used to study the involvement of CREB in sleep and waking (L. Graves, et al., unpublished observations cited in Hendricks, 2001). Findings support a conserved role of CREB in maintaining normal levels of wakefulness, independent of changes in circadian period. Additional studies to discover whether the rest-related role of CREB is, indeed, related to optimizing plasticity will continue to enhance understanding of the role of this signaling pathway in complex behaviors (Hendricks, 2001).

Because of the array of mutants and transgenics available in Drosophila, it was possible to show that baseline rest duration is inversely related to each component of the classic cAMP-PKA-dCREB2 signaling pathway. However, CREB and the CRE binding site are responsive to signals in addition to the cAMP-PKA pathway, and cAMP signaling has targets in addition to CREB. Multiple signaling pathways may well be involved in the many functions of CREB, and specifically in the rebound response to rest/sleep deprivation. CREB is a complex gene, even in Drosophila, with functions depending on the cellular milieu and developmental stage of the organism. The mammalian CREB/CREM family of transcription factors is critical for development, addiction, neural growth and survival, antidepressant effects, long-term memory consolidation and stress responses. Despite the potential for complex regulation of CREB activity, the data suggests that cAMP and PKA are particularly important in regulating CREB during the rest-activity cycle. An additional link between the circadian system and rest-activity behavior is a gene (the Drosophila NF-1 homolog) previously linked to learning and to cAMP signaling in both Drosophila and mammals, that modulates both the circadian rest-activity cycle and CREB activity. Based on the similarities in the effects of CREB mutations on rest in Drosophila and sleep in mammals, a role for NF-1 may similarly be conserved. Similar to the involvement of fruit fly genetics in identifying the molecular basis of vertebrate circadian rhythms, studies of the molecular mechanisms of Drosophila rest should help to focus studies of mammalian sleep (Hendricks, 2001).

Induction of cAMP response element-binding protein-dependent medium-term memory by appetitive gustatory reinforcement in Drosophila larvae: larval memory depends on both amnesiac and CREB

Drosophila has been successfully used as a model animal for the study of the genetic and molecular mechanisms of learning and memory. Although most of the Drosophila learning studies have used the adult fly, the relative complexity of its neural network hinders cellular and molecular studies at high resolution. In contrast, the Drosophila larva has a simple brain with uniquely identifiable neural networks, providing an opportunity of an attractive alternative system for elucidation of underlying mechanisms involved in learning and memory. This paper describes a novel paradigm of larval associative learning with a single odor and a positive gustatory reinforcer, sucrose. Mutant analyses have suggested importance of cAMP signaling and potassium channel activities in larval learning as has been demonstrated with the adult fly. Intriguingly, larval memory produced by the appetitive conditioning lasts medium term and depends on both amnesiac and cAMP response element-binding protein (CREB). A significant part of memory was disrupted at very early phase by CREB blockade without affecting immediate learning performance. Moreover, synaptic output of larval mushroom body neurons is required for retrieval but not for acquisition and retention of the larval memory, including the CREB-dependent component (Honjo, 2005).

The larval olfactory system is significantly simpler than the adult system with only 21 odorant receptor neurons. To find chemicals that are suitable for larval learning assays, 30 odorants were examined for naive larval chemotactic behavior and they were classified into four groups based on their attractiveness. 19 moderate attractants were examined for their effectiveness on larval appetitive olfactory conditioning. Larvae were exposed to an odor for 30 min in association with 1 M sucrose spread on agar. After conditioning, larvae were gently rinsed with distilled water to remove sucrose and tested for olfactory response on the test plate. For 10 of the 19 odorants, animals that received the odor with 1 M sucrose showed enhanced migration to the conditioned odor with significantly higher RI than control larvae, which had been exposed to the same odor but in conjunction with distilled water. Among the odorants examined, linalool (LIN), Pentyl acetate (PA), and gamma-valerolactone (GVA), which gave the largest RI increments in LIN/SUC conditioning, were chosen (Honjo, 2005).

To examine whether the increase of response index after conditioning is attributable to associative learning, a set of control experiments were performed. Significant response index increase was observed only when larvae were trained with LIN in association with SUC (LIN/SUC); response index did not change significantly from naive larvae when larvae are trained with LIN in association with distilled water (LIN) or sucrose alone (SUC). Notably, neither LIN nor sucrose alone resulted in habituation of larval olfactory responses compared with naive animals. Similar results were obtained with PA, except that conditioning with PA in association with distilled water led to slight desensitization. In contrast, conditioning with GVA in association with distilled water led to strong desensitization. However, the associative conditioning with GVA/SUC overcame the suppression (Honjo, 2005).

It was then asked whether the enhancement of larval response requires simultaneous exposure to both the odor and the reinforcer. As a temporal dissociation control, larvae were successively exposed first to sucrose and then to LIN or vise versa. Whereas simultaneous exposure to both LIN and sucrose (conditioning 1) resulted in enhanced olfactory response, the dissociation control, in which larvae were first exposed to sucrose and then to LIN, led to no enhancement compared with the odor alone control (conditioning 2). The requirement of temporal association between odor exposure and sucrose reinforcement was further confirmed in another set of dissociation controls. Exposure to LIN (conditioning 5) led to slightly higher larval response than conditioning 2, which seems a nonassociative effect caused by the delay attributable to the 30 min mock treatment (for delayed nonassociative effects). Nonetheless, simultaneous exposure to LIN and 1 M sucrose (conditioning 4) led to additional response index increment reproducing associative odor learning. In contrast, separate exposures to LIN and then 1 M SUC (conditioning 6) failed to do so (Honjo, 2005).

It was next asked whether the increased larval olfactory response was specific to the exposed odor. To address this question, larval olfactory responses were tested using odorants other than the one used for conditioning. When larvae were trained with LIN/SUC, PA/SUC, or GVA/SUC, only those trained with LIN/SUC showed significant response index increment in the olfactory test with LIN. Similarly, only larvae trained with PA/SUC showed significant response index increment in the olfactory test with PA. These results thus demonstrate that the enhanced larval response with sucrose is specific to the conditioned odor and suggest that Drosophila larvae discriminate the three odors despite their limited olfactory system (Honjo, 2005).

Whereas the above data emphasizes the importance of sucrose as a positive reinforcer, it is not clear whether response index stimulation is attributable to gustatory stimuli or attributable to higher osmotic pressure of 1 M sucrose than that of distilled water. To clarify this point, larvae were trained with LIN in association with 1 MD-sorbitol, a sugar that is tasteless to the flies. Conditioning with LIN in association with D-sorbitol failed to stimulate larval response index compared with the control, in which larvae were exposed to LIN in association with distilled water (Honjo, 2005).

Most studies on Drosophila associative learning have used reciprocal and symmetrical experimental paradigms with two odors. In contrast, the paradigm here uses only a single odor for conditioning and test. Consequently, this asymmetric nature calls for parallel controls to rule out nonassociative learning such as habituation and sensitization. Nonetheless, the paradigm resulted in significant learning only by the associative conditioning, in which both an odor and sucrose were simultaneously presented to larvae; enhanced larval olfactory response was specific to the odor paired with sucrose, excluding nonassociative sensitization to a broad range of odors. Conversely, it should be noted that strong desensitization was observed for certain odors such as GVA. Even with LIN, which showed no desensitization in immediate learning, delayed nonassociative effects on larval olfactory response were detected, emphasizing the importance of odor choice and careful data interpretation (Honjo, 2005).

Because different sets of larvae are used for control experiments for the stimuli involved, reproducibility of larval responses is critical to the paradigm. At this point, select odorants were used for screened for larval olfactory learning. Thus, of 30 chemicals, several odorants were chosen that produced significant response index increment with sucrose. Many odorants, such as 1-octanol and 4-methylcyclohexanol, which have been used in adult studies, failed to produce significant response index increment. The fact that larvae and adult flies exhibit different olfactory responses also highlights the importance of odorant choice for larval experiments (Honjo, 2005).

Despite its asymmetric design, several points are notable with regard to this paradigm: (1) the simple experimental design minimizes stress on larvae, which could affect learning performance; (2) the paradigm generates medium term memory (MTM) that lasts up to 3 h; (3) the paradigm is free from odor discriminative task. Because the larval olfactory system is considerably simpler than the adult system, simultaneous discrimination of different odors could complicate animal responses, although other studies have used two-odor paradigms; (4) because only a single odor is applied to larvae during training, the simple design of this paradigm may be of use in imaging of neural representation of the conditioned odor in the brain during learning and memory (Honjo, 2005).

Adult flies with amn mutations show a reduction in immediate memory as well as a more profound reduction in MTM. In this paradigm, amn larvae show reduced but significant immediate learning/memory. In the adult brain, the AMN peptide is expressed in dorsal paired medial (DPM) neurons that are situated medially to MBs and ramify throughout the MB lobes. Little is known about the network of the DPM neurons and the AMN expression pattern in the larval brain (Honjo, 2005).

Studies with Aplysia, mice, and adult Drosophila flies show that CREB-dependent transcription is required for cellular events underlying LTM. These studies have shown that CREB functions as a conserved molecular switch for LTM, which is thought to be induced several hours after training. Moreover, intervals between trainings or stimulations are generally required to produce CREB-dependent long-term effects (Honjo, 2005).

The larval CREB-dependent memory is stable for only medium term. Moreover, this paradigm continuously exposes larvae to an odor and sucrose during training, a condition similar to massed training of adult flies. Intriguingly, CREB is recruited shortly after learning in larvae; a significant portion of 30 min memory was disrupted by the CREB blocker, whereas immediate learning was not. If the larval MTM is induced after STM as in the adult fly, this very early CREB requirement might imply fast transition of memory phases. Alternatively, the CREB-dependent memory might also be generated independently. Intriguingly, it has been proposed that CREB can be activated independent of STM in long-term synaptic facilitation in Aplysia. In addition, although memory performance becomes undetectable in 3 h, the requirement of CREB activity suggests neural mechanisms that are in part shared with LTM in the adult fly. In fact, memory decay after CREB blockade is somewhat slower than in amn mutants. Furthermore, whereas CREB blocker has been shown to suppress 1 and 7 d memories, whether the blockade has more immediate effects is not known, leaving the possibility that CREB could be recruited early in the adult fly as well. Notably, memory performance in the adult fly tends to be higher with spaced training than with massed training already at several hours (Honjo, 2005).

Biochemically, CREB is activated by phosphorylation in response to diverged extra cellular stimuli. Among them, the protein kinase A (PKA) plays a central role in phosphorylation of CREB1-a, the catalytic subunit. Because the larval memory is completely disrupted in dnc and rut mutants, the cAMP-PKA pathway might be involved in the early activation of CREB in larvae. Alternatively, intracellular pathways other than PKA could also be recruited to mediate CREB activation. Intriguingly, increase of intracellular cAMP is known to activate mitogen-activated protein kinase in Aplysia, which in turn phosphorylates CREB2-b, the regulatory subunit, allowing transcriptional activation by the catalytic CREB isoform in the nuclei (Honjo, 2005).

The adult MBs are highly complex structures with three sets of lobes, each of which might participates in different memory traces. In contrast, the larval MBs exhibit a remarkably simple projection pattern with only a single set of lobes. In addition, recent studies have revealed straightforward organization of the larval olfactory system with only 21 olfactory receptor neurons targeting the 21 antennal lobe glomeruli, from which projection neurons target the larval MB calyx that consists of ~28 glomeruli (Honjo, 2005).

The finding that larval MB output is essential for memory retrieval discloses functional importance of the larval MBs and directly demonstrates anatomical commonality of memory networks between the larval and adult brains. Furthermore, the results that larval MB output is not required for memory acquisition and retention suggest that larval olfactory memory is localized upstream of larval MB synapses, in either MB neurons themselves or upstream circuits such as antennal lobes. Combined with the recent advances in functional neural imaging, the simple and identifiable neural network of the larval olfactory system will help further elucidation of the cellular basis of learning and memory in the brain (Honjo, 2005).

Waking experience affects sleep need in Drosophila: Experience-dependent changes in sleep need require dopaminergic modulation, cAMP signaling, and a particular subset of long-term memory genes

Sleep is a vital, evolutionarily conserved phenomenon, whose function is unclear. Although mounting evidence supports a role for sleep in the consolidation of memories, until now, a molecular connection between sleep, plasticity, and memory formation has been difficult to demonstrate. Drosophila as a model to investigate this relation; the intensity and/or complexity of prior social experience stably modifies sleep need and architecture. Furthermore, this experience-dependent plasticity in sleep need is subserved by the dopaminergic and adenosine 3',5'-monophosphate signaling pathways and a particular subset of 17 long-term memory genes (Ganguly-Fitzgerald, 2006).

Sleep is critical for survival, as observed in the human, mouse, and fruit fly, and yet, its function remains unclear. Although studies suggest that sleep may play a role in the processing of information acquired while awake, a direct molecular link between waking experience, plasticity, and sleep has not been demonstrated. Advantage was taken of Drosophila genetics and the behavioral and physiological similarities between fruit fly and mammalian sleep to investigate the molecular connection between experience, sleep, and memory (Ganguly-Fitzgerald, 2006).

Drosophila is uniquely suited for exploring the relation between sleep and plasticity for at least two reasons. (1) Fruit flies sleep. This is evidenced by consolidated periods of quiescence associated with reduced responsiveness to external stimuli and homeostatic regulation -- the increased need for sleep that follows sleep deprivation. (2) Drosophila has been successfully used to elucidate conserved mechanisms of plasticity. For example, exposure to enriched environments, including the social environment, affects the number of synapses and the size of regions involved in information processing in vertebrates and Drosophila. In the fruit fly, these structural changes occur in response to experiential information received within a week of emergence from pupal cases. Although brain plasticity is not limited to this period, the first week of emergence does coincide with the development of complex behaviors in Drosophila, including sleep. Hence, daytime sleep, which accounts for about 40% of total sleep in adults, is highest immediately after eclosion and stabilizes to adult levels 4 days after emergence (Ganguly-Fitzgerald, 2006).

To assess the impact of waking experience during this period of brain and behavioral development, individuals from the wild-type C-S strain were exposed to either social enrichment or impoverishment immediately at eclosion and were tested individually for sleep 5 days later. Socially enriched individuals (E), exposed to a group of 30 or more males and females (1:1 sex ratio) before being tested, slept significantly more than their socially impoverished (I) siblings, who were housed individually. This difference in sleep [DeltaSleep (E)] was restricted to daytime sleep. Socially enriched individuals consolidated their daytime sleep into longer bouts of ~60 min compared with their isolated siblings, who slept in 15-min bouts. In contrast, nighttime sleep was unaffected by prior social experience, corresponding with observations that daytime sleep is more sensitive to sex, age, genotype, and environment, when compared with nighttime sleep. This effect of social experience on sleep persisted over a period of days. Moreover, it was a stable phenotype: When socially enriched, longer-sleeping individuals and socially impoverished, shorter-sleeping siblings were sleep-deprived for 24 hours, they defended their respective predeprivation baseline sleep quotas by returning to these levels after a normal homeostatic response (Ganguly-Fitzgerald, 2006).

Experience-dependent modifications in sleep have long been observed in humans, rats, mice, and cats. But what is the nature of the experiential information that modifies sleep need in genetically identical Drosophila? Differences in sleep need in socially enriched and socially impoverished individuals were not a function of the space to which they were exposed -- flies reared in 2-cc tubes slept the same as those reared in 40-cc vials. Neither did it arise out of differences in reproductive state or sexual activity between the two groups: Socially impoverished mated and virgin individuals slept the same, as did socially enriched individuals from mixed-sex or single-sex groups. Further, differences in sleep were not a reflection of differences in overall activity (measured as infrared beam breaks) between the two groups. Although social context can reset biological rhythms, mutations in clock (Clkjerk), timeless (tim01), and cycle (cyc01) disrupt circadian rhythms but had no effect on experience-dependent responses in sleep need (Ganguly-Fitzgerald, 2006).

Because social interaction requires sensory input, fly strains that were selectively impaired in vision, olfaction, and hearing were evaluated . Blind norpA homozygotes failed to display a response in sleep to waking experience: Sleep need in norpA mutants did not increase after exposure to social enrichment. In contrast, norpA/+ heterozygotes with restored visual acuity slept more when previously socially enriched. Attenuating visual signals by rearing wild-type (C-S) flies in darkness also abolished the effect of waking experience on sleep. Compromising the sense of smell while retaining visual acuity also blocked experience-dependent changes in sleep need: Socially enriched smellblind1 mutants slept the same as their impoverished siblings. As confirmation, neurons carrying olfactory input to the brain were specifically silenced [Or83b-Gal4/UAS-TNT, and it was observed that sleep in these flies was also not affected by prior waking experience. Auditory cues, however, did not affect the relation between experience and sleep. Finally, sleep need in individual Drosophila increased with the size of the social group to which they were previously exposed. Socially isolated flies slept the least, whereas those exposed to social groups of 4, 10, 20, 60, and 100 (1:1 sex ratio) showed proportionately increased daytime sleep need. When rendered blind, however, flies did not display this relation between sleep need and the intensity of prior social interactions (Ganguly-Fitzgerald, 2006).

If sensory stimulation received during a critical period of juvenile development directs the maturation of the adult sleep homeostat, then subsequent environmental exposure should not affect adult sleep time and consolidation. Alternatively, if experience-dependent modifications in sleep are a reflection of ongoing plastic processes, this phenomenon would persist in the adult. It was observed that sleep in flies was modified by their most recent social experience regardless of juvenile experience. Shorter sleeping socially impoverished adults became longer sleepers when exposed to social enrichment before being assayed. Conversely, longer sleeping socially enriched flies became shorter sleepers after exposure to a period of social isolation. Moreover, repeated switching of exposure between the two social environments consistently modified sleep, reflecting an individual's most recent experience (Ganguly-Fitzgerald, 2006).

An estimation of neurotransmitter levels in whole brains revealed that short-sleeping, socially impoverished individuals contained one-third as much dopamine as their longer-sleeping, socially stimulated isogenic siblings. Silencing or ablating the dopaminergic circuit in the brain [TH-Gal4/UAS-TNT and TH-Gal4/UAS-Rpr specifically abolished response to social impoverishment in individuals that were reared in social enrichment. Similar results were obtained when endogenous dopamine levels were aberrantly increased, by disrupting the monoamine catabolic enzyme, arylalkylamine N-acetyltransferase, in Datlo mutants. Hence, abnormal up- or down-regulation of the dopaminergic system prevented behavioral plasticity in longer sleeping, socially enriched individuals when switched to social impoverishment (Ganguly-Fitzgerald, 2006).

The observation that dopaminergic transmission affects experience-dependent plasticity in sleep need is particularly compelling, given its role as a modulator of memory. Mutations in 49 genes implicated in various stages of learning and memory were screened to assess their impact on experience-dependent changes in sleep need. Of these, only mutations in short- and long-term memory genes affected experience-dependent plasticity in sleep need. Mutations in dunce (dnc1) and rutabaga (rut2080) have opposite effects on intracellular levels of adenosine 3',5'-monophosphate (cAMP), but are both correlated with short-term memory loss. In dnc1 mutants, waking experience had no impact on subsequent sleep need. This effect was partially rescued in dnc1/+ heterozygotes, but complete rescue was only achieved when a fully functional dunce transgene was introduced into the null mutant background. rut2080, however, selectively abolished the ability of socially enriched adults to demonstrate decreases in sleep after exposure to social impoverishment, which was reminiscent of aberrant dopaminergic modulation. Similarly, of the long-term memory genes screened, 17 (~40%) specifically disrupted the change in sleep need in socially enriched adults after exposure to social impoverishment. For example, overexpression of the Drosophila CREB gene repressor, dCREB-b, resulted in socially enriched flies that continued to be longer sleepers even after exposure to social impoverishment. As a control, overexpression of the dCREB-a activator yielded wild-type phenotypic read out. It is noteworthy that not all long-term memory mutants had a disrupted relation between experience and sleep. Instead, the particular subset of genes identified, only half of which are expressed in the mushroom bodies, may specifically contribute to pathways that underlie sleep-dependent consolidation of memories (Ganguly-Fitzgerald, 2006).

Finally, to assess the correlation between sleep and memory, male flies trained for a courtship conditioning task that generated long-term memories were measured for sleep after training. Males whose courtship attempts are thwarted by nonreceptive, recently mated females or by males expressing aphrodisiac pheromones form long-term associative memories as evidenced by subsequently reduced courtship of a receptive virgin female. Trained males that formed long-term memories slept significantly more than their untrained siblings and wake controls (ones that were sleep-deprived while the experimental flies were being trained). Exposure to a virgin female did not alter sleep need. As before, this increase in sleep was associated with longer daytime sleep bouts in trained individuals compared with controls. Further, sleep deprivation for 4 hours immediately after training abolished training-induced changes in sleep-bout duration, as well as courtship memory. Although these results are intriguing, invertebrate memory is particularly sensitive to extinction by mechanical perturbations. However, gentle handling that ensured wakefulness, but not mechanical stimulation, immediately following training, also abolished subsequent courtship memory. Furthermore, sleep deprivation per se did not affect the formation of long-term memory: Trained flies that were allowed to sleep unperturbed for 24 hours and then subjected to 4 hours of sleep deprivation retained courtship memory (Ganguly-Fitzgerald, 2006).

In summary, this study has demonstrate a rapid and dynamic relation between prior social experience and sleep need in Drosophila. In particular, experience-dependent changes in sleep need require dopaminergic modulation, cAMP signaling, and a particular subset of long-term memory genes, supporting the hypothesis that sleep and neuronal activity may be inexorably intertwined. These observations are compelling given two recent studies have demonstrating a central role of the mushroom bodies in sleep regulation and emphasize the importance of establishing Drosophila as a model system to investigate the molecular pathways underlying sleep and plasticity (Ganguly-Fitzgerald, 2006).

A new genetic model of activity-induced Ras signaling dependent pre-synaptic plasticity in Drosophila

Techniques to induce activity-dependent neuronal plasticity in vivo allow the underlying signaling pathways to be studied in their biological context. This study demonstrates activity-induced plasticity at neuromuscular synapses of Drosophila double mutant for comatose (an NSF mutant) and Kum (Calcium ATPase at 60A: a SERCA mutant), and presents an analysis of the underlying signaling pathways. comt; Kum (CK) double mutants exhibit increased locomotor activity under normal culture conditions, concomitant with a larger neuromuscular junction synapse and stably elevated evoked transmitter release. The observed enhancements of synaptic size and transmitter release in CK mutants are completely abrogated by: a) reduced activity of motor neurons; b) attenuation of the Ras/ERK signaling cascade; or c) inhibition of the transcription factors Fos and CREB. All of which restrict synaptic properties to near wild type levels. Together, these results document neural activity-dependent plasticity of motor synapses in CK animals that requires Ras/ERK signaling and normal transcriptional activity of Fos and CREB. Further, novel in vivo reporters of neuronal Ras activation and Fos transcription also confirm increased signaling through a Ras/AP-1 pathway in motor neurons of CK animals, consistent with results from the genetic experiments. Thus, this study: a) provides a robust system in which to study activity-induced synaptic plasticity in vivo; b) establishes a causal link between neural activity, Ras signaling, transcriptional regulation and pre-synaptic plasticity in glutamatergic motor neurons of Drosophila larvae; and c) presents novel, genetically encoded reporters for Ras and AP-1 dependent signaling pathways in Drosophila (Freeman, 2010).

This study describes a new model for activity-dependent pre-synaptic plasticity in Drosophila. In the double mutant combination of comt and Kum, sustained elevation of neural activity (potentially including seizure-like motor neuron firing under normal rearing conditions) results in the expansion of motor synapses with a concomitant increase in transmitter release. These synaptic changes are mediated by the Ras/ERK signaling cascade and the activity of at least two key transcription factors, CREB and Fos. In vivo reporter assays also directly demonstrate Ras activation and enhanced transcription of Fos in the nervous system. CK is the only genetic model of synaptic plasticity in Drosophila in which pre-synaptic plasticity has been correlated with the Ras/ERK signaling cascade. This result is especially relevant given the wide conservation of the Ras/ERK signaling cascade in plasticity and recent demonstrations of the involvement of this signaling cascade in learning behavior in flies (Godenschwege, 2004; Moressis, 2009). Significant insights into Ras mediated regulation of both synapse growth and transmitter release are also presented (Freeman, 2010).

Non-invasive methods to manipulate neural activity in select neurons continue to be an important experimental target in plasticity research. In Drosophila, combinations of the eag and Shaker potassium channel mutants have long been used to chronically alter neural activity and study downstream cellular events. In recent years, transgenic expression of modified Shaker channels has also been generated and used to alter excitability in both neurons and muscles. However, the CK model of activity-dependent plasticity was developed since in synaptic changes in CK were consistently more robust than eag Sh and core plasticity-related signaling components were activated in a predictable manner in CK mutants. Another advantage with CK is the option of acutely inducing seizures as has been used to identify activity-regulated genes. CK thus combines advantages of both eag Sh and seizure mutants, and as is shown in this study, leads to an activity-dependent increase in synaptic size and transmitter release. It is believed that this model will prove highly beneficial to the large community of researchers who investigate synaptic plasticity in Drosophila. The utility of more recent techniques (such as the ChannelRhodopsin or the newly reported temperature sensitive TrpA1 channel transgenes) to induce neural activity-dependent synaptic plasticity at Drosophila motor synapses has not been tested yet and it will be interesting to see if these afford greater experimental flexibility in the future (Freeman, 2010).

Signal transduction through the Ras cascade has been shown to affect both dendritic and pre-synaptic plasticity in invertebrate and vertebrate model systems. In mammalian neurons, Ras signaling has been linked to hippocampal slice LTP, changes in dendritic spine architecture and plasticity of cultured neurons. In this context, Ras signaling has been shown to impinge on downstream MAP kinase signaling, thus implicating a canonical signaling module already established as a mediator of long-term plasticity in vertebrates. In Drosophila, expression of a mutant constitutively active Ras that is predicted to selectively target ERK leads to synapse expansion and increased localized phosphorylation of ERK at pre-synaptic terminals. In light of these observations, tests were performed to see if Ras signaling os necessary and sufficient for synaptic plasticity in CK. The results suggest that synaptic changes in CK are driven by stimulated Ras/ERK signaling in Drosophila motor neurons, and these can be replicated by directly enhancing Ras signaling in these cells. Furthermore Ras activation was found to be sufficient to cause stable elevation in pre-synaptic transmitter release. Finally, evidence is provided to show that synaptic effects of Ras activation require the function of both Fos and CREB in motor neurons. The consistency of signaling events in CK with those observed in mammalian preparations makes this a more useful and generally applicable genetic model of synaptic plasticity (Freeman, 2010).

In vivo reporters of neural activity have been difficult to design but offer better experimental resolution and flexibility over standard immuno-histochemical or RNA in situ methods to detect changes in gene expression in the brain. Thus, a good reporter permits increased temporal and spatial resolution, the option of live imaging (for fluorescent reporters) and in the case of transcriptional reporters, better understanding of cis-regulatory elements that control activity-dependent gene expression. This paper describes two genetically encoded reporters with utility clearly beyond the current study; a Raf based reporter to detect Ras activation in neurons and an enhancer based reporter to detect transcription of Fos (Freeman, 2010).

The Ras binding domain of Raf has been used previously to detect Ras expression in yeast, mammalian cell lines, and recently in hippocampal neuron dendrites. This study used a similar strategy to model the reporter using the conserved Ras binding domain and the cysteine-rich domain (RBD + CRD) from Drosophila Raf, under the reasonable assumption that this would provide sensitive reporter activity in neurons. This is the first time that a Ras reporter has been utilized in an intact metazoan organism to measure changes in endogenous Ras activity. In addition to confirming Ras activation in CK brains, it is expected that this reporter will find widespread use in tracing Ras activation in multiple tissues through development and in response to signaling changes in the entire organism. Since the reporter is based on the GAL4-UAS system, it can be expressed in tissues of choice, limiting reporter activity to regions of interest. Indeed, the experiments with the eye-antennal imaginal disc illustrate the utility of this reporter in identifying regions of activated Ras signaling during eye development (Freeman, 2010).

The Fos transcriptional reporter is one of the very few activity-regulated reporters in existence in Drosophila and it should find broad acceptance as a tool to map neural circuits in the fly brain that show activity-dependent plasticity. The reporter believed to be reasonably accurate since it is expressed in expected tissue domains (embryonic leading edge cells, for instance), and also co-localizes extensively with anti-Fos staining in the larval brain. There are several recognizable transcription factor binding motifs that can be detected in this 5 kb region of DNA (including binding sites for CREB, Fos, Mef2 and c/EBP). Which of these transcription factors regulate activity-dependent Fos expression from this enhancer is currently unknown. However, future experiments that dissect functional elements in this large enhancer region are expected to refine and identify these regulatory elements. Such studies are likely to lead the way in the development of a new generation of neural activity reporters in the brain (Freeman, 2010).

Regulation of energy stores and feeding by neuronal and peripheral CREB activity in Drosophila

The cAMP-responsive transcription factor CREB functions in adipose tissue and liver to regulate glycogen and lipid metabolism in mammals. While Drosophila has a homolog of mammalian CREB, dCREB2, its role in energy metabolism is not fully understood. Using tissue-specific expression of a dominant-negative form of CREB (DN-CREB), this stud examined the effect of blocking CREB activity in neurons and in the fat body, the primary energy storage depot that functions as adipose tissue and the liver in flies, regulating energy balance, stress resistance and feeding behavior. It was found that disruption of CREB function in neurons reduces glycogen and lipid stores and increases sensitivity to starvation. Expression of DN-CREB in the fat body also reduces glycogen levels, while it does not affect starvation sensitivity, presumably due to increased lipid levels in these flies. Interestingly, blocking CREB activity in the fat body increased food intake. These flies do not show a significant change in overall body size, suggesting that disruption of CREB activity in the fat body causes an obese-like phenotype. Using a transgenic CRE-luciferase reporter, it was further demonstrated that disruption of the adipokinetic hormone receptor, which is functionally related to mammalian glucagon and beta-adrenergic signaling, in the fat body reduces CRE-mediated transcription in flies. This study demonstrates that CREB activity in either neuronal or peripheral tissues regulates energy balance in Drosophila, and that the key signaling pathway regulating CREB activity in peripheral tissue is evolutionarily conserved (Iijima, 2009).

This study provides in vivo evidence that both neuronal and peripheral CREB activities are involved in the regulation of energy balance in flies. Blocking CREB activity in neurons causes reductions in both glycogen and lipid stores and a higher sensitivity to starvation stress. In contrast, while disruption of CREB function in the fat body also reduces glycogen levels, it increases lipid stores, and does not affect starvation sensitivity. Since there was no significant change in overall body size in these flies, disruption of CREB activity in the fat body causes an obese-like phenotype. These results also indicate that CREB activity can both increase and reduce lipid stores in flies depending on its site of action. Recently, two distinct populations of Drosophila brain neurons that regulate fat deposition were identified in Drosophila (Al-Anzi, 2009). It will be interesting to determine in which neurons CREB functions to regulate energy metabolism in flies (Iijima, 2009).

In a recent study, TORC-mediated CREB activity in neurons was shown to positively regulate glycogen and lipid stores in flies. This is based on results showing that expression of TORC in neurons rescues the starvation sensitivity of TORC mutant flies. In addition, expression of TORC in neurons partially rescues the lower energy stores of these mutants. While supporting the conclusions of this study with respect to the role of neuronal CREB activity, the current results also provide evidence that CREB in the fat body plays roles in energy balance. Moreover, in contrast to the normal feeding behavior of a TORC mutant, it was found that blocking CREB activity in the fat body increases food intake. Thus, disruption of CREB functions has a broader impact on energy metabolism and feeding behavior than the loss of TORC. It is likely that not all CREB functions depend on TORC. In support of this, although a TORC null mutant is viable and fertile, CREB mutants are lethal (Iijima, 2009).

This study found that adipokinetic hormone (AKH/AKHR) signaling in the fat body, which is thought to be functionally related to glucagon/glucogon receptor signaling in the mammalian liver, positively regulates CRE-mediated transcription. In the mammalian liver, CREB activates the gluconeogenic program following a glucagon stimulus. Recent studies reported that promoting AKH signaling in the fat body significantly reduces, while loss of AKHR function modestly increases, glycogen levels in flies, presumably through AKH/AKHR-mediated carbohydrate catabolism in the fat body (Gronke, 2007; Bharucha, 2008). However, this study found that blocking CREB activity in the fat body significantly reduces glycogen levels, which would seem to contradict the proposed role of AKH/AKHR in mediating carbohydrate catabolism in the fat body. One possibility is that CREB activity in the fat body regulates multiple aspects of glucose/glycogen metabolism in addition to the AKH/AKHR-mediated pathway, and that blocking all CREB functions in the fat body reduces total glycogen levels as a net effect. In fact, significant CREB activity was remaining in AKHR mutant flies, suggesting that other signaling pathways might contribute to the activation of CREB activity in the fat body. Further studies will be required to delineate the role of CREB activity in the fat body in carbohydrate metabolism and its relationship with the AKH signaling pathway (Iijima, 2009).

This study found that blocking CREB activity in the fat body increased lipid stores. AKH/AKHR is also thought to be functionally related to β-adrenergic signaling in mammalian adipose tissue, which activates protein kinase A (PKA) and stimulates lipolysis by phosphorylating hormone-sensitive lipase and perilipin. In Drosophila, the promotion of AKH signaling in the fat body reduces lipid levels, whereas loss of AKHR function has the opposite effect; this is partly ascribed to altered activity in lipocatabolic systems. In addition, AKH signaling has been shown to repress the lipogenesis pathway in various insects. Interestingly, blocking CREB activity in mammalian liver causes excessive fat accumulation, resulting in 'fatty liver' through overactivation of liposynthesis. Future analysis will unravel whether CREB activity in the fat body represses liposynthesis and/or promotes lipid catabolism under the control of AKH/AKHR signaling (Iijima, 2009).

In summary, these results demonstrate that CREB is involved in both central and peripheral regulation of energy balance and feeding behavior in Drosophila. Future studies of CREB in flies hold great promise for revealing the mechanisms underlying energy balance and feeding behavior. Such studies will likely contribute to understanding of human metabolic disorders (Iijima, 2009).

Light-induced structural and functional plasticity in Drosophila larval visual system

How to build and maintain a reliable yet flexible circuit is a fundamental question in neurobiology. The nervous system has the capacity for undergoing modifications to adapt to the changing environment while maintaining its stability through compensatory mechanisms, such as synaptic homeostasis. This study describes findings in the Drosophila larval visual system, where the variation of sensory inputs induces substantial structural plasticity in dendritic arbors of the postsynaptic neuron and concomitant changes to its physiological output. Furthermore, a genetic analysis has identified the cyclic adenosine monophosphate (cAMP) pathway and a previously uncharacterized cell surface molecule as critical components in regulating experience-dependent modification of the postsynaptic dendrite morphology in Drosophila (Yuan, 2011).

Proper functions of neuronal circuits rely on their fidelity, as well as plasticity, in responding to experience or changing environment, including the Hebbian form of plasticity, such as long-term potentiation, and the homeostatic plasticity important for stabilizing the circuit. Activity-dependent modification of neuronal circuits helps to establish and refine the nervous system and provides the cellular correlate for cognitive functions, such as learning and memory. Multiple studies have examined synaptic strength regulation by neuronal activity, whereas to what extent and how the dendritic morphology may be modified by neuronal activity remain open questions (Yuan, 2011).

The model organism Drosophila melanogaster has facilitated genetic studies of nervous system development and remodeling. Notwithstanding the relatively stereotyped circuitry, flies exhibit experience-induced alterations in neuronal structures and behaviors such as learning and memory). In a study of experience-dependent modifications of the Drosophila larval CNS, it has been found that different light exposures dramatically altered dendritic arbors of ventral lateral neurons [LN(v)s], which are postsynaptic to the photoreceptors. Unlike the visual activity-induced dendrite growth in Xenopus optic tectum, extending the light exposure of Drosophila larvae reduced the LN(v)s' dendrite length and functional output, a homeostatic plasticity for compensatory adaptation to alterations in sensory inputs. It was further shown that the cyclic adenosine monophosphate (cAMP) pathway and an immunoglobulin domain-containing cell surface protein, CG3624, are critical for this sensory experience-induced structural plasticity in Drosophila CNS (Yuan, 2011).

In Drosophila larvae, Bolwig's organ (BO) senses light, and its likely postsynaptic targets are LN(v)s. As the major circadian pacemaker, LN(v)s are important for the entrainment to environmental light-dark cycles and larval light avoidance behavior. In the larval brain, Bolwig's nerve (BN), the axonal projection from BO, terminates in an area overlapping the dendritic field of LN(v)s. Using the FRT-FLP system [in which DNA sequences flanked by flippase recognition targets (FRT) are snipped out by flippase (FLP)] along with three-dimensional (3D) tracing, the dendritic arbor of individual LN(v) neurons were labeled and analyzed. Then potential synaptic connections were demonstrated between BN and LN(v)s using the GRASP [green fluorescent protein (GFP) reconstitution across synaptic partners] technique to drive expression of one-half of the split GFP in the BN by means of Gal4/UAS and expression of the other half of the split GFP in LN(v)s via LexA/LexAop. The proximity of putative synaptic connections between BN and LN(v)s' dendrites reconstituted GFP fluorescence for photoreceptors expressing either rhodopsin 5 (Rh5) or rhodopsin 6 (Rh6) in BO, which suggested that both groups of photoreceptors may have synaptic connections with LN(v)s (Yuan, 2011).

To test whether LN(v)s can be activated by BN inputs through light stimulation, calcium imaging was performed using GCaMP3 transgenic flies with the larval brain-eye preparation, which included BO, BN, developing eye disks, the larval brain, and ventral nerve cord. Because BO senses blue and green light, the confocal laser at 488 nm (blue) and 543 nm (green) were used to stimulate these larval photoreceptors. LN(v)s' axon terminals displayed a relatively stable baseline of GCaMP3 fluorescence and, upon light stimulation, yielded large calcium responses, which increased with the greater intensity and longer duration of the light pulses (Yuan, 2011).

Recent studies suggest that Cryptochrome (CRY) in adult large LN(v)s senses light and elicits neuronal firing. In larvae, however, severing BN abolished light-induced calcium responses in LN(v)s. The loss-of-function mutation of NorpA (no-receptor-potential A), encoding a phospholipase C crucial for phototransduction, also eliminated these calcium responses, which indicated that light-elicited responses in LN(v)s are generated via phototransduction in larval photoreceptors rather than as a direct response to light by LN(v)s (Yuan, 2011).

In animals with BO genetically ablated, the dendritic field of LN(v) is absent. To test whether BO is required for LN(v)s' dendrite maintenance, the expression of cell death genes rpr and hid was induced in BO after synapse formation, and the LN(v) dendrite length was also found to be greatly reduced. Whereas physical contacts with BN or growth-promoting factors released from presynaptic axons could be important for LN(v)s' dendrite maintenance, it is also possible that synaptic activity from BN promotes LN(v) dendrite growth, as suggested by previous studies. To explore the latter scenario, newly hatched larvae were provided with different visual experiences through various light regimes—including the standard 12 hours of light and 12 hours of dark daily cycle (LD); constant darkness (DD) for sensory deprivation; constant light (LL) for enhanced light input; 16-hour light and 8-hour dark cycle, mimicking a long day; and 8-hour light and 16-hour dark cycle, mimicking a short day. The dendrite morphology of LN(v)s of late third instar larvae was examined. Whereas different light exposure had no detectable effects on larval developmental timing, increasing light exposure reduced the total dendrite length of individual LN(v) neurons, with the longest dendrite in constant darkness and the shortest dendrite length in constant light condition. Thus, not only is the LN(v) dendrite dependent on the presence of presynaptic nerve fibers, its length is modified by the sensory experience in a compensatory fashion, whereby an increase in sensory inputs causes a reduction in the dendrite length and vice versa (Yuan, 2011).

Whereas adult LN(v)s alter their axon terminal structures in a circadian cycle-controlled fashion, no difference was found in dendrite morphology of LN(v)s from larvae collected at four different time points around the clock, which indicated that circadian regulation is not involved in the light-induced modification of LN(v) dendrites. Under regular light-dark conditions, LN(v) dendrites expanded as the larval brain size increased from the second to the third instar stage. However, the dendrite length of the LL group increased at a significantly slower rate than the DD group. It thus appears that light exposure retards the growth of LN(v) dendrites throughout the larval development (Yuan, 2011).

To test the contribution of different light-sensing pathways, loss-of-function mutations of Cry (cry01) or NorpA (norpA36) and of double mutants lacking both Rh5 and Rh6 (rh52;rh61) were examined. Although wild-type and cry01 larvae displayed differences in their dendrite length when exposed to constant darkness versus constant light, such light-induced changes were absent in the rh52;rh61 double mutant and the norpA36 mutant. Thus, similar to the calcium response to light, light-induced modification of LN(v) dendritic structure requires visual transduction mediated by rhodopsin and NorpA in BO but not Cry function in LN(v)s (Yuan, 2011).

To manipulate the level of synaptic activity, the BO excitability was weither increased by expressing the heat-activated Drosophila transient-receptor-potential A1 (dTrpA) channel, or transmitter release from BN was reduced through a temperature-sensitive form of the dominant-negative dynamin, Shibirets (Shits). These manipulations eliminated light-induced modification of LN(v) dendrites at 29°C. Reducing BO activity by means of Shits caused dendrite expansion, as if the animal detected no light, whereas increasing BO activity by means of the dTrpA channel resulted in reduction of LN(v) dendrites, a process reminiscent of constant light exposure (Yuan, 2011).

Whether intrinsic LN(v) neuronal activity drives modification of its dendrite morphology was further tested by expression of either the sodium channel NaChBac to increase excitability or the potassium channel Kir2.1 to reduce excitability. LN(v)s expressing Kir2.1 showed reduced or no calcium responses upon light stimulation. In contrast, LN(v)s expressing NaChBac displayed numerous peaks in GCaMP3 signals in the presence or absence of light stimulation, indicative of elevated spontaneous activities. Upon examining LN(v) dendrites, it was found that neuronal excitability of the LN(v) was inversely proportional to its dendrite length (Yuan, 2011).

These results obtained using genetic approaches agreed with findings in experiments with different environmental light conditions. They suggested that LN(v)'s dendritic structures are modified according to its neuronal activity, which varies with light-induced synaptic inputs (Yuan, 2011).

To test whether synaptic contacts of BN on LN(v)s are modified by light, synapses formed by BN with EGFP (enhanced green fluorescent protein)-tagged Cacophony (Cac-EGFP) were marked, because Cacophony is a calcium channel localized at presynaptic terminals and its distribution correlates with the number of synapses. Close association was found of Cac-EGFP-expressing structures with LN(v)s' dendritic arbors. Compared with regular light-dark conditions, constant darkness increased, whereas constant light reduced, the total intensity of Cac-EGFP, which suggested that light modified not only dendritic arbors of LN(v)s but also the number of synaptic contacts impinging on LN(v) dendrites (Yuan, 2011).

Next, using calcium imaging, whether there are light-induced functional modifications of LN(v)s was examined. Increased light exposure caused LN(v)s to be less responsive. Conversely, sensory deprivation in constant darkness increased LN(v)s' sensitivity to light. Thus, in contrast to stable synaptic responses observed in synaptic homeostasis, light-induced responses of central neurons postsynaptic to photoreceptors in the Drosophila larval visual circuit have a dynamic range, modifiable by sensory experiences and positively correlated to the dendrite length (Yuan, 2011).

In dunce1, a loss-of-function mutant of the fly homolog of 3'5'-cyclic nucleotide phosphodiesterase, the LN(v)s' dendrite length was comparable among LD, LL, and DD groups. Reducing dunce gene expression specifically in LN(v)s through RNA interference (dncIR) resulted in a similar indifference of LN(v)s' dendrite size to the light exposure, which implicated a cell-autonomous action of dunce in LN(v) neurons (Yuan, 2011).

To explore the possibility that the elevated cAMP level caused by the dunce mutation interfered with dendrite plasticity, tests were performed for the involvement of downstream components of the cAMP pathway, including the catalytic subunit of protein kinase A (PKAmc), which up-regulates cAMP signaling, and a dominant-negative form of the cAMP response element-binding protein (CREBdn), which inhibits cAMP-induced transcription activation. Expression of either transgene specifically in LN(v)s obliterated their ability to adjust dendrite length under different light-dark conditions. Calcium imaging further revealed that the expression of PKAmc or CREBdn eliminated changes of LN(v)s' light responses produced by different light-dark conditions. Thus, the cAMP pathway regulates both structural and functional plasticity of LN(v)s (Yuan, 2011).

The screen for mutants with defective LN(v) dendritic plasticity also identified babos-1, a mutant with a P-element insertion near the transcriptional start site of CG3624, a previously uncharacterized immunoglobulin domain-containing cell surface protein. The LN(v) dendrite length of babos-1 mutant larvae was comparable to controls in LD and LL but has no compensatory increase in DD. Similar phenotypes were found in larvae expressing an RNAi transgene targeting CG3624 in LN(v)s. Moreover, flies carrying a hypomorphic allele of CG3624, CG3624[KG05061], also showed defective light-induced dendritic plasticity, which was fully rescued by expressing the UAS-CG3624 transgene specifically in LN(v)s. Thus, the function of this immunoglobulin domain-containing protein in LN(v)s is important for the dendrite expansion in constant darkness (Yuan, 2011).

Bioinformatic analyses suggest that CG3624 is a cell surface protein containing an N-terminal signal peptide, extracellular immunoglobulin domains followed by a transmembrane helix, and a short C-terminal cytoplasmic tail. CG3624 is widely expressed in the nervous system throughout development. Its specific requirement for the adjustment of LN(v)s' dendrite length in constant darkness suggests that elevation or reduction of sensory inputs likely invokes separate mechanisms for compensatory modifications of central neuronal dendrites (Yuan, 2011).

A functioning nervous system must have the capacity for adaptive modifications while maintaining circuit stability. This study of the Drosophila larval visual circuit reveals large-scale, bidirectional structural adaptations in dendritic arbors invoked by different sensory exposure. Whereas the circuit remains functional with modified outputs, this type of homeostatic compensation may modify larval light sensitivity according to its exposure during development and could facilitate adaption of fly larvae toward altered light conditions, such as seasonal changes. The observations also suggest shared molecular machinery between homeostasis and the Hebbian plasticity with respect to the cAMP pathway and indicate the feasibility of genetic studies of experience-dependent neuronal plasticity in Drosophila (Yuan, 2011).


REFERENCES (part 1/2)

Ahn, S., et al. (1998). A dominant-negative inhibitor of CREB reveals that it is a general mediator of stimulus-dependent transcription of c-fos. Mol. Cell. Biol. 18(2): 967-977. PubMed Citation: 9447994

Ahn, S., Ginty, D. G. and Linden, D. J. (1999). A late phase of cerebellar long-term depression requires activation of CaMKIV and CREB. Neuron 23: 559-568. PubMed Citation: 10433267

Akimaru, H., et al. (1997a). Drosophila CBP is a co-activator of cubitus interruptus in hedgehog signalling. Nature 386: 735-738. PubMed Citation: 9109493

Akimaru, H., Hou, D. X. and Ishii, S. (1997b). Drosophila CBP is required for dorsal-dependent twist gene expression. Nat. Genet. 17(2): 211-214. PubMed Citation: 9326945

Al-Anzi, B., et al. (2009). Obesity-blocking neurons in Drosophila. Neuron 63: 329-341. PubMed Citation: 19679073

Alberts, A. S., et al. (1994). Recombinant cyclic AMP response element binding protein (CREB) phosphorylated on Ser-133 is transcriptionally active upon its introduction into fibroblast nuclei. J. Biol. Chem. 269: 7623-30. PubMed Citation: 8125987

Arnould, T., et al. (2002). CREB activation induced by mitochondrial dysfunction is a new signaling pathway that impairs cell proliferation. EMBO J. 21: 53-63. 11782425

Aso, Y., Hattori, D., Yu, Y., Johnston, R. M., Iyer, N. A., Ngo, T. T., Dionne, H., Abbott, L. F., Axel, R., Tanimoto, H. and Rubin, G. M. (2014). The neuronal architecture of the mushroom body provides a logic for associative learning. Elife 3: e04577. PubMed ID: 25535793

Bailey, C.H. and Kandel, E.R. (1994). Structural changes underlying long-term memory storage in Aplasia: a molecular prospective. Semin. Neurosci. 6: 35-44

Barco, A., Alarcon, J. M. and Kandel1, E. R. (2002). Expression of constitutively active CREB protein facilitates the late phase of long-term potentiation by enhancing synaptic capture. Cell 108: 689-703. 11893339

Barco, A., et al. (2005). Gene expression profiling of facilitated L-LTP in VP16-CREB mice reveals that BDNF is critical for the maintenance of LTP and its synaptic capture. Neuron 48: 123-137. 16202713

Bartsch, D., et al. (1995). Aplysia CREB2 represses long-term facilitation: relief of repression converts transient facilitation into long-term functional and structural change. Cell 83(6): 979-92. PubMed Citation: 8521521

Bartsch, D., et al. (2000). Enhancement of memory-related long-term facilitation by ApAF, a novel transcription factor that acts downstream from both CREB1 and CREB2. Cell 103: 595-608. PubMed Citation: 11106730

Benito, E. and Barco, A. (2010). CREB's control of intrinsic and synaptic plasticity: implications for CREB-dependent memory models. Trends Neurosci 33: 230-240. PubMed ID: 20223527

Bharucha, K. N., Tarr, P. and Zipursky, S. L. (2008). A glucagon-like endocrine pathway in Drosophila modulates both lipid and carbohydrate homeostasis. J. Exp. Biol. 211: 3103-3110. PubMed Citation: 18805809

Belvin, M. P., Zhou, H. and Yin, J. C. (1999). The Drosophila dCREB2 gene affects the circadian clock. Neuron 22(4): 777-87. PubMed Citation: 10230797

Birchenall-Roberts. M. C., et al. (1995). Nuclear localization of v-Abl leads to complex formation with cyclic AMP response element (CRE)-binding protein and transactivation through CRE motifs. Mol. Cell. Biol. (11): 6088-6099. PubMed Citation: 7565761

Bito, H., Deisseroth, K., and Tsien, R. W. (1996). CREB phosphorylation and dephosphorylation: A Ca++ and stimulus duration-dependent switch for hippocampal gene expression. Cell 87: 1203-14. PubMed Citation: 8980227

Bonni, A., et al. (1999). Cell survival promoted by the Ras-MAPK signaling pathway by transcription-dependent and -independent mechanisms. Science 286: 1358-1362. PubMed Citation: 10558990

Casadio, A., et al. (1999). A transient, neuron-wide form of CREB-mediated long-term facilitation can be stabilized at specific synapses by local protein synthesis. Cell 99: 221-237. PubMed Citation: 10535740

Chawla, S., et al. (1998). CBP: A signal-regulated transcriptional coactivator controlled by nuclear calcium and CaM Kinase IV. Science 281(5382): 1505-1509. PubMed Citation: 9727976

Chen, A., et al. (2003). Inducible enhancement of memory storage and synaptic plasticity in transgenic mice expressing an inhibitor of ATF4 (CREB-2) and C/EBP proteins, Neuron 39: 655-669. Medline abstract: 12925279

Chen, C. C., Wu, J. K., Lin, H. W., Pai, T. P., Fu, T. F., Wu, C. L., Tully, T. and Chiang, A. S. (2012). Visualizing long-term memory formation in two neurons of the Drosophila brain. Science 335: 678-685. PubMed ID: 22323813

Chen, X. and Ganetzky, B. (2012). A neuropeptide signaling pathway regulates synaptic growth in Drosophila. J Cell Biol 196: 529-543. PubMed ID: 22331845

Chen YC, Chen HJ, Tseng WC, Hsu JM, Huang TT, Chen CH, Pan CL. (2016). A C. elegans thermosensory circuit regulates longevity through crh-1/CREB-dependent flp-6 neuropeptide signaling. J Dev Cell 39(2):209-223. PubMed ID: 27720609

Cheng, H. Y., et al. (2007). microRNA modulation of circadian-clock period and entrainment. Neuron 54(5): 813-29. Medline abstract: 17553428

Cho, Y. H., et al. (1998). Abnormal hippocampal spatial representations in alphaCaMKIIT286A and CREBalphaDelta- mice. Science 279(5352): 867-869. PubMed Citation: 9452387

Clarke, N., et al. (1998). Epidermal growth factor induction of the c-jun promoter by a Rac pathway. Mol. Cell. Biol. 18(2): 1065-1073. PubMed Citation: 9448004

Conkright, B. D., et al. (2003). Genome-wide analysis of CREB target genes reveals a core promoter requirement for cAMP responsiveness. Molec. Cell 11: 1101-1108. 12718894

Costa-Mattioli, M., et al. (2005). Translational control of hippocampal synaptic plasticity and memory by the eIF2alpha kinase GCN2. Nature 436: 1166-1173. Medline abstract: 16121183

Costa-Mattioli, M., et al. (2007). eIF2α phosphorylation bidirectionally regulates the switch from short- to long-term synaptic plasticity and memory. Cell 129: 195-206. Medline abstract: 17418795

Crino, P., et al. (1998). Presence and phosphorylation of transcription factors in developing dendrites. Proc. Natl. Acad. Sci. 95: 2313-2318. PubMed Citation: 9482882

Das, S., et al. (1997). NMDA and D1 receptors regulate the phosphorylation of CREB and the induction of c-fos in striatal neurons in primary culture. Synapse 25(3): 227-233. PubMed Citation: 9068120

Dash, P. K., Tian, L. M. and Moore, A. N. (1998). Sequestration of cAMP response element-binding proteins by transcription factor decoys causes collateral elaboration of regenerating Aplysia motor neuron axons. Proc. Natl. Acad. Sci. 95(14): 8339-8344. PubMed Citation: 9653188

Davis, G. W., Schuster, C. M. and Goodman, C. S. (1996). Genetic dissection of structural and functional components of synaptic plasticity. III. CREB is necessary for presynaptic functional plasticity. Neuron 17: 669-679. PubMed Citation: 8893024

Davis, R. L., et al. (1996). Physiology and biochemistry of Drosophila learning mutants. Physiol. Rev. 76: 299-317. PubMed Citation:

Deak, M., et al. (1998). Mitogen- and stress-activated protein kinase-1 (MSK1) is directly activated by MAPK and SAPK2/p38, and may mediate activation of CREB. EMBO J. 17: 4426-4441. PubMed Citation: 8618959

De Cesare, D., et al. (1998). Rsk-2 activity is necessary for epidermal growth factor-induced phosphorylation of CREB protein and transcription of c-fos gene. Proc. Natl. Acad. Sci. 95(21): 12202-7. PubMed Citation: 9770464

Deisseroth, K., Bito, H. and Tsien, R. W. (1996). Signaling from synapse to nucleus: postsynaptic CREB phosphorylation during multiple forms of hippocampal synaptic plasticity. Neuron 16: 89-101. PubMed Citation: 8562094

Dockendorff, T. C., et al. (2002). Drosophila lacking dfmr1 activity show defects in circadian output and fail to maintain courtship interest. Neuron 34: 973-984. 12086644

Dooley, K. A., Bennett, M. K., Osborne, T. F. (1999). A critical role for cAMP response element-binding protein (CREB) as a co-activator in sterol-regulated transcription of 3-hydroxy-3-methylglutaryl coenzyme A synthase promoter. J. Biol. Chem. 274(9): 5285-91. PubMed Citation: 10026135

Drain, P., Folkers, E. and Quinn, W. G. (1991). cAMP-dependent protein kinase and the disruption of learning in transgenic flies. Neuron 6: 71-82. PubMed Citation: 1702651

Edelman, D. B., Meech, R. and Jones, F. S. (2000). The homeodomain protein Barx2 contains activator and repressor domains and interacts with members of the CREB family. J. Biol. Chem. 275: 21737-21745. 10781615

Ernst, P., et al. (2001). MLL and CREB bind cooperatively to the nuclear coactivator Creb-binding protein. Mol. Cell. Bio. 21: 2249-2258. 11259575

Eresh, S., et al. (1997). A CREB-binding site as a target for decapentaplegic signalling during Drosophila endoderm induction. EMBO J. 16: 2014-22. PubMed Citation: 9155027

Feliciello, A., et al. (1997). A-kinase anchor protein 75 increases the rate and magnitude of cAMP signaling to the nucleus. Curr. Biol. 7(12): 1011-1014. PubMed Citation: 9382844

Fimia, G. M, De Cesare, D. and Sassone-Corsi, P. (1999). CBP-independent activation of CREM and CREB by the LIM-only protein ACT. Nature 398(6723): 165-9. PubMed Citation: 10086359

Finkbeiner, S., et al. (1997). CREB: a major mediator of neuronal neurotrophin responses. Neuron 19(5): 1031-1047. PubMed Citation: 9390517

Freeman, A., et al. (2010). A new genetic model of activity-induced Ras signaling dependent pre-synaptic plasticity in Drosophila. Brain Res. 1326: 15-29. PubMed Citation: 20193670

Frey, U. and Morris, R. G. (1997) Synaptic tagging and long-term potentiation. Nature, 385: 533-536. 9020359

Frey, U. and Morris, R. G. (1998) Weak before strong: dissociating synaptic tagging and plasticity-factor accounts of late-LTP. Neuropharmacology 37: 545-552. 9704995

Fropf, R., Tubon, T. C., Jr. and Yin, J. C. (2013). Nuclear gating of a Drosophila dCREB2 activator is involved in memory formation. Neurobiol Learn Mem 106C: 258-267. PubMed ID: 24076014

Fropf, R., Zhang, J., Tanenhaus, A. K., Fropf, W. J., Siefkes, E. and Yin, J. C. (2014). Time of day influences memory formation and dCREB2 proteins in Drosophila. Front Syst Neurosci 8: 43. PubMed ID: 24744705

Fujita, H., Fujii, R., Aratani, S., Amano, T., Fukamizu, A. and Nakajima, T. (2003). Antithetic effects of MBD2a on gene regulation. Mol. Cell. Biol. 23(8): 2645-57. 12665568

Ganguly-Fitzgerald, I., Donlea, J. and Shaw, P. J. (2006). Waking experience affects sleep need in Drosophila. Science 313(5794): 1775-81. PubMed Citation: 16990546

Gao, Y., et al. (2004). Activated CREB is sufficient to overcome inhibitors in myelin and promote spinal axon regeneration in vivo. Neuron 44: 609-621. 15541310

Gass, P., et al. (1998). Deficits in memory tasks of mice with CREB mutations depend on gene dosage. Learn Mem. 5(4-5): 274-88. PubMed Citation: 10454354

Giebler, H. A., et al. (1997). Anchoring of CREB binding protein to the human T-cell leukemia virus type 1 promoter: a molecular mechanism of Tax transactivation. Mol. Cell. Biol. 17(9): 5156-5164. PubMed Citation: 9271393

Giese, K., et al. (1995). Assembly and function of a TCR alpha enhancer complex is dependent on LEF-1-induced DNA bending and multiple protein-protein interactions. Genes Dev. 9(8): 995-1008. PubMed Citation: 7774816

Ginty, D. D., et al. (1993) Regulation of CREB phosphorylation in the suprachiasmatic nucleus by light and a circadian clock. Science 260: 238-241. PubMed Citation: 8097062

Ginty, D. D., Bonni, A. and Greenberg, M. E. (1994). Nerve growth factor activates a Ras-dependent protein kinase that stimulates c-fos transcription via phosphorylation of CREB. Cell 77: 713-25. PubMed Citation: 8205620

Gronke, S., et al. (2007). Dual lipolytic control of body fat storage and mobilization in Drosophila. PLoS Biol 5: e137. PubMed Citation: 17488184

Godenschwege, T. A., et al. (2004). Flies lacking all synapsins are unexpectedly healthy but are impaired in complex behaviour. Eur. J. Neurosci. 20: 611-622. PubMed Citation: 15255973

Gu, W., Shi, X. L. and Roeder, R. G. (1997a). Synergistic activation of transcription by CBP and p53. Nature 387(6635): 819-823. PubMed Citation: 9194564

Guo, B., et al. (1997). ATF1 and CREB trans-activate a cell cycle regulated histone H4 gene at a distal nuclear matrix associated promoter element. Biochemistry 36(47): 14447-14455. PubMed Citation: 9398163

Hagiwara, M., et al. (1993). Coupling of hormonal stimulation and transcription via the cyclic AMP-responsive factor CREB is rate limited by nuclear entry of protein kinase A. Mol Cell Biol 13: 4852-9. PubMed Citation: 8336722

Halder, R., Hennion, M., Vidal, R. O., Shomroni, O., Rahman, R. U., Rajput, A., Centeno, T. P., van Bebber, F., Capece, V., Garcia Vizcaino, J. C., Schuetz, A. L., Burkhardt, S., Benito, E., Navarro Sala, M., Javan, S. B., Haass, C., Schmid, B., Fischer, A. and Bonn, S. (2016). DNA methylation changes in plasticity genes accompany the formation and maintenance of memory. Nat Neurosci 19(1): 102-110. PubMed ID: 26656643

Hegde, A. N., et al. (1997). Ubiquitin C-terminal hydrolase is an immediate-early gene essential for long-term facilitation in Aplysia. Cell 89: 115-126. PubMed Citation: 9094720

Hendricks, J. C., et al. (2001). A non-circadian role for cAMP signaling and CREB activity in Drosophila rest homeostasis. Nat. Neurosci. 4: 1108-1115. 11687816

Hirano, Y., Masuda, T., Naganos, S., Matsuno, M., Ueno, K., Miyashita, T., Horiuchi, J. and Saitoe, M. (2013). Fasting launches CRTC to facilitate long-term memory formation in Drosophila. Science 339: 443-446. PubMed ID: 23349290

Hirano, Y., Ihara, K., Masuda, T., Yamamoto, T., Iwata, I., Takahashi, A., Awata, H., Nakamura, N., Takakura, M., Suzuki, Y., Horiuchi, J., Okuno, H. and Saitoe, M. (2016). Shifting transcriptional machinery is required for long-term memory maintenance and modification in Drosophila mushroom bodies. Nat Commun 7: 13471. PubMed ID: 27841260

Honjo, K. and Furukubo-Tokunaga, K. (2005). Induction of cAMP response element-binding protein-dependent medium-term memory by appetitive gustatory reinforcement in Drosophila larvae. J. Neurosci. 25(35): 7905-13. 16135747

Hou, D. X., Akimaru, H. and Ishii, S. (1997). Trans-activation by the Drosophila myb gene product requires a Drosophila homologue of CBP. FEBS Lett. 413(1): 60-64. PubMed Citation: 9287117

Huh, G. S., et al. (2000). Functional requirement for class I MHC in CNS development and plasticity. Science 290: 2155-2159. 11118151

Iijima, K., Zhao, L., Shenton, C. and Iijima-Ando, K. (2009). Regulation of energy stores and feeding by neuronal and peripheral CREB activity in Drosophila. PLoS One 4(12): e8498. PubMed Citation: 20041126

Imai, T., Suzuki, M. and Sakano, H. (2006). Odorant receptor-derived cAMP signals direct axonal targeting. Science 314(5799): 657-61. Medline abstract: 16990513

Impey, S., et al. (1996). Induction of CRE-mediated gene expression by stimuli that generate long-lasting LTP in area CA1 of the hippocampus. Neuron 16: 973-982. PubMed Citation: 8630255

Impey, S., et al. (1998a). Cross talk between ERK and PKA is required for Ca2+ stimulation of CREB-dependent transcription and ERK nuclear translocation. Neuron 21(4): 869-83. PubMed Citation: 9808472

Impey, S., et al. (1998b). Stimulation of cAMP response element (CRE)-mediated transcription during contextual learning. Nature Neurosci. 1(7): 595-601. PubMed Citation: 10196567

Impey, S., et al. (2004). Defining the CREB regulon. A genome-wide analysis of transcription factor regulatory regions. 119: 1041-1054. 15620361

Ishimoto, H., Sakai, T. and Kitamoto, T. (2009). Ecdysone signaling regulates the formation of long-term courtship memory in adult Drosophila melanogaster. Proc. Natl. Acad. Sci. 106: 6381-6386. PubMed Citation: 19342482

Jhala, U. S., et al. (2003). cAMP promotes pancreatic ß-cell survival via CREB-mediated induction of IRS2. Genes Dev. 17: 1575-1580. 12842910

Ji, R. R., et al. (1998). Specific agrin isoforms induce cAMP response element binding protein phosphorylation in hippocampal neurons. J. Neurosci. 18(23): 9695-9702. PubMed Citation: 9822730

Josselyn, S. A., et al. (2001). Long-term memory is facilitated by cAMP response element-binding protein overexpression in the amygdala. J. Neurosci. 21(7): 2404-2412. 11264314

Kaang, B. K., Kandel, E. R. and Grant, S. G. (1993). Activation of cAMP-responsive genes by stimuli that produce long-term facilitation in Aplysia sensory neurons. Neuron 10: 427-35. PubMed Citation: 8384857

Kawasaki, H., et al. (1998). p300 and ATF-2 are components of the DRF complex, which regulates retinoic acid- and E1A-mediated transcription of the c-jun gene in F9 cells. Genes Dev. 12: 233-245. PubMed Citation: 9436983

Kopp, M., Meissl, H. and Korf, H. W. (1997). The pituitary adenylate cyclase-activating polypeptide-induced phosphorylation of the transcription factor CREB (cAMP response element binding protein) in the rat suprachiasmatic nucleus is inhibited by melatonin. Neurosci Lett. 227(3): 145-148. PubMed Citation: 9185671

Kovacs, K. A., Steullet, P., Steinmann, M., Do, K. Q., Magistretti, P. J., Halfon, O. and Cardinaux, J. R. (2007). TORC1 is a calcium- and cAMP-sensitive coincidence detector involved in hippocampal long-term synaptic plasticity. Proc Natl Acad Sci U S A 104: 4700-4705. PubMed ID: 17360587

Kwok, R. P., et al. (1994). Nuclear protein CBP is a coactivator for the transcription factor CREB The transcription factor CREB binds to a DNA element known as the cAMP-regulated enhancer (CRE). Nature 370: 223-226. PubMed Citation: 7913207

Lamprecht, R., Hazvi, S. and Dudai, Y. (1997). cAMP response element-binding protein in the amygdala is required for long- but not short-term conditioned taste aversion memory. J. Neurosci. 17(21): 8443-8450. PubMed Citation: 9334416

Lane-Ladd, S. B., et al. (1997). CREB (cAMP response element-binding protein) in the locus coeruleus: biochemical, physiological, and behavioral evidence for a role in opiate dependence. J. Neurosci. 17(20): 7890-7901. PubMed Citation: 9315909

Ledo, F., Kremer, L., Mellström, B. and Naranjo, J. R. (2002). Ca2+-dependent block of CREB-CBP transcription by repressor DREAM. EMBO J. 21: 4583-4592. 12198160

Lee, P. T., Lin, G., Lin, W. W., Diao, F., White, B. H. and Bellen, H. J. (2018). A kinase-dependent feedforward loop affects CREBB stability and long term memory formation. Elife 7. PubMed ID: 29473541

Lee, W. P., Chiang, M. H., Chang, L. Y., Lee, J. Y., Tsai, Y. L., Chiu, T. H., Chiang, H. C., Fu, T. F., Wu, T. and Wu, C. L. (2020). Mushroom body subsets encode CREB2-dependent water-reward long-term memory in Drosophila. PLoS Genet 16(8): e1008963. PubMed ID: 32780743

Lenzmeier, B. A., Giebler, H. A. and Nyborg, J. K. (1998). Human T-cell leukemia virus type 1 Tax requires direct access to DNA for recruitment of CREB binding protein to the viral promoter. Mol. Cell. Biol. 18(2): 721-731. PubMed Citation: 9447968

Liang, G. and Hai, T. (1997). Characterization of Human Activating Transcription Factor 4, a Transcriptional Activator That Interacts with Multiple Domains of cAMP-responsive Element-binding Protein (CREB)-binding Protein. J. Biol. Chem. 272(38): 24088-24095. PubMed Citation: 9295363

Lin, H. W., Chen, C. C., de Belle, J. S., Tully, T. and Chiang, A. S. (2021). CREBA and CREBB in two identified neurons gate long-term memory formation in Drosophila. Proc Natl Acad Sci U S A 118(37). PubMed ID: 34507985

Lin, H. W., Chen, C. C., Jhang, R. Y., Chen, L., de Belle, J. S., Tully, T. and Chiang, A. S. (2022). CREBB repression of protein synthesis in mushroom body gates long-term memory formation in Drosophila. Proc Natl Acad Sci U S A 119(50): e2211308119. PubMed ID: 36469774

Liu, F.-C. and Graybiel, A. M. (1996). Spatiotemporal dynamics of CREB phosphorylation: Transient versus sustained phosphorylation in the developing striatum. Neuron 17: 1133-1144. PubMed Citation: 8982161

Liu, F. C. and Graybiel, A. M. (1998). Region-dependent dynamics of cAMP response element-binding protein phosphorylation in the basal ganglia. Proc. Natl. Acad. Sci. 95(8): 4708-4713. PubMed Citation: 9539803

Liu, R. Y., Cleary, L. J. and Byrne, J. H. (2011). The requirement for enhanced CREB1 expression in consolidation of long-term synaptic facilitation and long-term excitability in sensory neurons of Aplysia. J Neurosci 31: 6871-6879. PubMed ID: 21543617

Long, F., et al. (2001). The CREB family of activators is required for endochondral bone development. Development 128: 541-550. 11171337

Lonze, B. E., et al. (2002). Apoptosis, axonal growth defects, and degeneration of peripheral neurons in mice lacking CREB. Neuron 34: 371-385. 11988169

Lu, P. D., Harding, H. P. and Ron, D. (2004). Translation reinitiation at alternative open reading frames regulates gene expression in an integrated stress response. J. Cell Biol. 167: 27-33. Medline abstract: 15479734

Lutz, B., et al. (1999). Essential role of CREB family proteins during Xenopus embryogenesis Mech. Dev. 88: 55-66. PubMed Citation: 10525188

Ma, H., Groth, R. D., Cohen, S. M., Emery, J. F., Li, B., Hoedt, E., Zhang, G., Neubert, T. A. and Tsien, R. W. (2014). γCaMKII shuttles Ca(2+)/CaM to the nucleus to trigger CREB phosphorylation and gene expression. Cell 159: 281-294. PubMed ID: 25303525

Mair, W., et al. (2011). Lifespan extension induced by AMPK and calcineurin is mediated by CRTC-1 and CREB. Nature 470(7334): 404-8. PubMed Citation: 21331044

Marie, H., et al. (2005). Generation of silent synapses by acute in vivo expression of CaMKIV and CREB. Neuron 45: 741-752. 15748849

Maronde, E., et al. (1999). Transcription factors in neuroendocrine regulation: rhythmic changes in pCREB and ICER levels frame melatonin synthesis. J. Neurosci. 19(9): 3326-36. PubMed Citation: 10212292

Martin, K. C., et al. (1997). Synapse-specific, long-term facilitation of aplysia sensory to motor synapses: a function for local protein synthesis in memory storage. Cell 91(7): 927-938. PubMed Citation: 9428516

Martin, M. D., et al. (1998). Repeated pulses of serotonin required for long-term facilitation activate mitogen-activated protein kinase in sensory neurons of aplysia. Proc. Natl. Acad. Sci. 95(4): 1864-1869. PubMed Citation:

Matthews, R. P., et al. (1994). Calcium/calmodulin-dependent protein kinase types II and IV differentially regulate CREB-dependent gene expression. Mol. Cell. Biol. 14: 6107-6116. PubMed Citation: 8065343

Mayall, T. P., et al. (1997). Distinct roles for P-CREB and LEF-1 in TCRalpha enhancer assembly and activation on chromatin templates in vitro Genes Dev. 11: 887-899. PubMed Citation: 9106660

McNulty, S., et al. (1998) Stimuli which entrain the circadian clock of the neonatal Syrian hamster in vivo regulate the phosphorylation of the transcription factor CREB in the suprachiasmatic nucleus in vitro. Eur. J. Neurosci. 10: 1063-1072. PubMed Citation: 9753174

Miyashita, T., Kikuchi, E., Horiuchi, J. and Saitoe, M. (2018). Long-term memory engram cells are established by c-Fos/CREB transcriptional cycling. Cell Rep 25(10): 2716-2728. PubMed ID: 30517860

Monnier, D. and Loeffler, J. P. (1998). Pituitary adenylate cyclase-activating polypeptide stimulates proenkephalin gene transcription through AP1- and CREB-dependent mechanisms. DNA Cell Biol. 17(2): 151-159. PubMed Citation: 9502431

Moressis, A. et al. (2009). A dual role for the adaptor protein DRK in Drosophila olfactory learning and memory. J. Neurosci. 29: 2611-2625. PubMed Citation: 19244537

Murphy, D. D. and Segal, M. (1997). Morphological plasticity of dendritic spines in central neurons is mediated by activation of cAMP response element binding protein. Proc. Natl. Acad. Sci. 94(4): 1482-1487. PubMed Citation: 9037079

Miyashita, T., Oda, Y., Horiuchi, J., Yin, J. C., Morimoto, T. and Saitoe, M. (2012). Mg2+ block of Drosophila NMDA receptors is required for long-term memory formation and CREB-dependent gene expression. Neuron 74(5): 887-98. PubMed Citation: 22681692

References continued: part 2/2


CrebB-17A: Biological Overview | Evolutionary Homologs | Regulation | Developmental Biology | Effects of Mutation

Home page: The Interactive Fly © 1997 Thomas B. Brody, Ph.D.

The Interactive Fly resides on the
Society for Developmental Biology's Web server.