14-3-3zeta/leonardo


DEVELOPMENTAL BIOLOGY

Embryonic

The D14-3-3 gene (14-3-3zeta or Leonardo) expresses 1.0-, 1.9- and 2.9-kb mRNAs, all of which show differential expression patterns. While the 2.9-kb mRNA is expressed only in the head, the other two mRNAs are found both in the head and body. Compared to the 1.9- and 2.9-kb mRNAs, the 1.0-kb mRNA is more abundant in the ovary and is probably maternally inherited. The 1.9-kb mRNA is the most predominant species in the embryo; its level peaks between 6-15 h of embryogenesis. The D14-3-3 gene is predominantly expressed in the ventral nerve cord of the embryo, and in the neural tissues of the head (Swanson, 1992).

Larval

In addition to a strong 14-3-3zeta expression in the central nervous system, an enrichment of the transcript is seen in the region posterior to the progressing morphogenic furrow of the developing eye imaginal disc. 14-3-3zeta (Leonardo) is expressed in most, if not all cells of the eye disc. Strong antibody staining levels are found in the region posterior to the morphogenetic furrow where cells undergo neuronal induction and differentiation as photoreceptors. Within these cells, the distribution of protein appears to be concentrated apically, showing a subcellular pattern similar to the distribution of proteins that act upstream of Raf, such as Boss, Sevenless, EGF receptor, Drk, Sos and Dos (Kockel, 1997)

Adult

Antisense probes produce remarkably intense signal in mushroom body cells. Lower levels of hybridization are observed in cells of the subesophageal ganglion, optic lobes, antennal lobes, and the central brain. The gene is also expressed in cells of the thoracic ganglia, nurse cells, and maturing oocytes of female flies (Sokoulakis, 1996).

Antibody preferentially decorates the neuropil of the mushroom bodies. The antigen is detected in the dendritic projections (calyces), the cytoplasm of the parikarya, and the axonal projections that form the peduncle and lobes of the mushroom body neurons. In addition, the antibody decorates the ellipsoid body, a neuropil structure of the central complex and a group of cell bodies residing just anterior to the "heel" of the mushroom bodies. These cells appear to be the ring neurons that project to the ellipsoid body. The antibody also decorates neuropil and cell bodies in the antennal lobes, albeit with lower intensity. In sagittal sections, modest staining is observed in thoracic ganglia and throughout the cytoplasm of nurse cells and oocytes (Sokoulakis, 1996).

Effects of Mutation or Deletion

Immunological localization of the Leonardo protein shows that it is expressed at synaptic connections and enriched in presynaptic boutons of the neuromuscular junction (NMJ). Null leonardo mutants die as mature embryos. Electrophysiological assays of the mutant NMJ demonstrate that basal synaptic transmission is reduced by 30% and that transmission amplitude, fidelity, and fatigue resistance properties are reduced at elevated stimulation frequencies and in low external [Ca2+]. Moreover, transmission augmentation and post-tetanic potentiation (PTP) are disrupted in the mutant. These results suggest that Leonardo plays a role in the regulation of synaptic vesicle dynamics, a function which may underlie synaptic modulation properties enabling learning (Broadie, 1997).

14-3-3 proteins have been shown to interact with Raf-1 and cause its activation when overexpressed. However, their precise role in Raf-1 activation is still enigmatic, as they are ubiquitously present in cells and found to associate with Raf-1 in vivo regardless of Raf's activation state. The function of the Drosophila 14-3-3 gene leonardo (leo) has been analyzed in the Torso (Tor) receptor tyrosine kinase (RTK) pathway. In the syncytial blastoderm embryo, activation of Tor triggers the Ras/Raf/MEK pathway that controls the transcription of tailless (tll). In the absence of Tor, overexpression of leo is sufficient to activate tll expression. The effect of leo requires D-Raf and Ras1 activities but not KSR or DOS, two recently identified essential components of Drosophila RTK signaling pathways. Tor signaling is impaired in embryos derived from females lacking maternal expression of leo. It is proposed that binding to 14-3-3zeta/Leonardo by Raf is necessary but not sufficient for the activation of Raf and that overexpressed Drosophila 14-3-3zeta requires Ras1 to activate D-Raf (Li, 1997).

Studies of Drosophila and other insects have indicated an essential role for the mushroom bodies in learning and memory. The leonardo gene encodes a Drosophila protein highly homologous to the vertebrate 14-3-3 zeta isoform. The gene is expressed abundantly and preferentially in mushroom body neurons. Mutant alleles that reduce Leonardo protein levels in the mushroom bodies significantly decrease the capacity for olfactory learning, but do not affect sensory modalities or brain neuroanatomy that are requisite for conditioning. These results establish a biological role for 14-3-3 proteins in mushroom body­mediated learning and memory processes, and suggest that proteins known to interact with them, such as RAF-1 or other protein kinases, may also have this biological function (Skoulakis, 1996).

Leonardo exhibits similar involvement in the Raf/Ras pathway. Clones of mutant leonardo show a loss of photoreceptors. Ommatidia lack outer as well as inner photoreceptors. This phenotype is reminiscent of clones homozygous for Drosophila ras or raf hypomorphic alleles. Artificial activation of Raf rescues the nonviability causes by leonardo mutation and permits photoreceptor development. It is concluded that leonardo acts downstream of Ras and upstream of Raf in the signaling pathway that controls cell proliferation in the Drosophila eye imaginal disc (Kockel, 1997).

The reduction of leonardo gene dose appears to worsen the phenotype of 14-3-3eta mutants, suggesting that the different 14-3-3 proteins are partially redundant. However, reduction of leonardo gene does not detectably modify the rough eye phenotype caused by activated Ras1 expression (Chang, 1997).

Members of the ubiquitous 14-3-3 family of proteins are abundantly expressed in metazoan neurons. The Drosophila 14-3-3zeta gene leonardo is preferentially expressed in adult mushroom bodies, centers of insect learning and memory. Mutants exhibit defects in olfactory learning and memory and physiological neuroplasticity at the neuromuscular junction. Because strong mutations in this gene are lethal, the nature of the defects that precipitate the learning and memory deficit and the role of the two protein isoforms encoded by leonardo in these processes were investigated. The behavioral deficit in the mutants is not caused by aberrant development, Leonardo protein is acutely required for learning and memory, and both protein isoforms can function equivalently in embryonic development and behavioral neuroplasticity (Philip, 2001).

The leonardo gene encodes three size classes of transcripts attributable to use of alternative promoters and three polyadenylation sites. Alternative splicing of exon 6 or 6' into the mRNA results in two protein isoforms (LeoI and LeoII) that differ by five amino acids. Because exons 6 and 6' are similar in size, alternative inclusion into the mRNA does not contribute to size heterogeneity. RT-PCR was used to determine the spatial and temporal expression of mRNAs that contain exon 6 (leoI) and 6' (leoII) (Philip, 2001).

Both leoI and leoII transcripts are present in embryos before activation of the zygotic genome, suggesting that they are deposited in the oocytes maternally. Exclusive presence of leoII transcripts in stage 10-12 embryos indicates preferential splicing of exon 6' into the mRNA, which may underlie a specific contribution of LeoII to early development. In contrast, both leoI and leoII transcripts are found in late embryos and all larval stages. In adult animals, although both isoforms are present in heads and abdomens, leoI is absent from the thorax (Philip, 2001).

To determine whether head tissues that require leo function exhibit differential isoform expression, flies carrying the eyes-absent mutation and wild-type animals were subjected to mushroom body ablation with hydroxyurea. Lack of eye tissues does not eliminate one of the isoforms differentially, but leoII is specifically absent from the brains of mushroom body-ablated animals. The results indicate that leoII transcripts are specific to the mushroom bodies, whereas although leoI may be present in these neurons, it is more broadly expressed in the brain. Outside the mushroom bodies, Leo protein is preferentially distributed in the ellipsoid body neurons of the central complex. Because ellipsoid body neurons are not ablated by hydroxyurea and retain Leo immunoreactivity, they must contain only leoI transcripts. Presence of D14-3-3epsilon in all tissues and stages tested suggests a broad role in basic cellular functions, and possible colocalization with Leo isoforms may result in heterodimer formation. Together, the differential expression of the two leonardo mRNAs in embryos and adult tissues suggests functional differences between the two Leo protein isoforms. Therefore, a functional investigation of potential differences between LeoI and LeoII isoforms was necessary before experiments aimed at rescuing the learning-memory deficit of leo mutations (Philip, 2001).

To investigate potential functional differences of the putative Leo isoforms, conditional rescue of lethality associated with strong leo alleles was attempted. The level of Leo protein induced in heads of rescued leoP1375 homozygotes was estimated. These homozygotes contained ~75%-80% the amount of Leo present in similarly treated wild-type animals. Because both LeoI and LeoII can support development to adulthood, under the conditions used the isoforms do not exhibit functional specificity. Furthermore, both LI and LII transgenes are highly inducible and allow manipulation of Leo levels in adult heads over a wide range, and animals thus obtained do not exhibit morphological defects. In addition, these experiments identified highly inducible leo transgenes and methods to obtain animals for behavioral testing (Philip, 2001).

The differential distribution of leo transcripts in adult heads suggested potentially differential roles for LeoI and LeoII in olfactory learning and memory. To investigate whether the behavioral deficit of leonardo viable alleles could be reversed by conditional induction of the leo transgenes, both transgenes were introduced into Df(1)yw67c23;leo23, (leo23) and Df(1)yw67c23;leoX1, (leoX1) flies. To ascertain that the transgenes remained inactive during development, all animals including the Df(1)yw67c23 (yw) control strains were raised at 18°C. Because leonardo expression in tissues other than the mushroom bodies and ellipsoid body appears normal in these alleles, quantitative Western blots were not used to monitor Leo levels in the heads of these animals. Transgene induction in animals raised at 18°C was achieved by two 30 min heat shocks delivered 6 hr apart, followed by a 5-6 hr rest period. Accumulation of LeoI and LeoII in the mushroom bodies of leo23;LI and leo23;LII animals after the rest period was monitored by immunohistochemistry using the anti-Leo antibody. Very low levels of Leo protein are present in the mushroom bodies of leo23;LI and leo23;LII animals. A significant increase of both protein isoforms in the mushroom bodies and ellipsoid body neurons was observed during induction of the respective transgenes, although final accumulation did not equal the amount of Leo in controls. Similar results were obtained with leoX1;LI and leoX1;LII animals. Moreover, lack of Leo during development did not precipitate neuroanatomical aberrations in the brains of mutant animals raised at 18°C (Philip, 2001).

Animals raised at 18°C and ones subjected to the induction and rest period were transferred to 23-24°C 2 hr before behavioral experiments. The growth conditions and temperature shift did not affect the ability of the mutants to perceive the stimuli used for olfactory conditioning compared with similarly treated controls. To further investigate their olfactory acuity, the performance of mutants and controls toward an attractive odor, geraniol, was measured using a modified olfactory trap assay. Although an attractive odor is not used in conditioning, this test is a good measure of olfactory acuity. Flies seek and navigate toward the source of an attractive odor, a more complex olfactory task than simple avoidance of an aversive odor. The performance of experimental animals was not significantly different from controls. Performance of the animals after olfactory conditioning was assessed immediately after training or 90 min later to investigate memory. The performance of leo23;LI, leo23;LII, leoX1;LI, and leoX1;LII animals exhibited a significant 30% decrement compared with controls both immediately and 90 min after training, similar to the decrement observed with leo23 and leoX1 animals raised under similar conditions. In contrast, learning and 90 min retention were not significantly different from controls during transgene induction before conditioning. The results suggest that LeoI and LeoII accumulation in the mushroom bodies after transgene induction fully restores the learning and memory deficit of leo23and leoX1 mutants. Interestingly, under the conditions used, both LeoI and LeoII isoforms appear equivalent in rescuing the behavioral deficit of the mutants. Collectively, the results indicate strongly that leonardo gene products are acutely required for mushroom body-dependent olfactory learning and memory (Philip, 2001).

Given the behavioral rescue of leo mutants, it was of interest to determine whether the learning and memory deficit exhibited by leo viable alleles represents the maximal contribution of Leo-mediated processes in mushroom body-dependent olfactory learning. The ability to obtain animals that harbor very low levels of Leo throughout their heads was used to address this question. The nearly complete lack of Leo throughout the adult brain did not result in neuroanatomical anomalies judged by histological and immunohistochemical analyses using multiple markers to examine the morphology of the mushroom bodies and central brain (Philip, 2001).

To determine whether lack of Leo throughout these rescued animals precipitated general sensory deficits, they were subjected to behavioral control tests. These leo mutants exhibited normal attraction to geraniol, avoidance of electric shock (US), and avoidance of both aversive odors (benzaldehyde and 3-octanol) used as CS. However, all rescued animals exhibit a 25%-30% decrement in olfactory learning. Significantly, the decrease in learning exhibited by the rescued lethal homozygotes and heteroallelics was similar in magnitude with that of leo23;LI animals. Therefore, near lack of Leo throughout the head does not reduce learning further than exhibited by leo23 animals, which lack Leo only in the mushroom bodies. This suggests that the leo23 and leoX1 mutations represent strong mutant alleles with respect to the behavioral phenotype. As with leo23;LI animals, the learning deficit of lethal homozygotes and heteroallelics was fully rescued to control levels by multiple inductions of either LI or LI/LII transgenes. To determine whether restoration of learning ability results from permanent changes attributable to the elevation of Leo, animals were kept at 18°C after induction and trained and tested along similarly treated and aged controls. Restoration of learning during transgene induction decayed back to mutant levels 60-70 hr later compared with age-matched control animals (Philip, 2001).

These results indicate that induction of Leo to levels sufficient to restore learning does not precipitate permanent changes but rather that the available amount of protein is acutely essential for this process. Furthermore, elevated Leo expression outside the mushroom bodies and ellipsoid body observed in controls and abrogated in the mutants does not appear essential for learning, the sensory inputs used in these experiments, or for the neuroanatomical integrity of the brain (Philip, 2001).

Thus, acute induction of either LI or LII transgenes completely restores learning and memory in leo23 and leoX1 mutant flies. The behavioral deficit in these animals is unlikely the result of sensory or developmental defects below the threshold of detection but rather are attributable to an acute requirement for Leo in learning-memory. This conclusion is further supported by the reversible rescue of the learning deficit exhibited by lethal homozygotes and heteroallelics. In contrast to leo23 and leoX1 mutants, which lack Leo in mushroom body and ellipsoid body neurons, minimally rescued animals contain a negligible amount of Leo throughout their heads. This lack of Leo in lethal homozygotes and heteroallelics should exaggerate putative developmental or sensory deficits present in leo23 and leoX1. However, neither sensory deficits nor anatomical aberrations were detectable in the later, despite the lack of transgene induction in larval or pupal stages. Therefore, either the 10%-15% residual Leo suffices for normal larval development and the reorganization of the brain at pupariation or Leo isoforms are not required for these processes. Because transgene induction and Leo accumulation restores the learning deficit of the lethal alleles to control levels, but this recovery is eliminated during the decay of the protein, Leo is acutely necessary for learning (Philip, 2001).

Involvement of 14-3-3 proteins in multiple cellular processes may be at least in part the result of multiple isotypes or isoforms present within one cell. This may be of particular importance in vertebrates in which nine isotypes exhibit primarily overlapping expression patterns, especially in neuronal tissues. Similarly, because Leo isoforms and D14-3-3epsilon appear to be at least partially overlapping, heterodimerization among the three 14-3-3 proteins is possible. In fact, genetic analysis suggested interactions between leonardo and D14-3-3epsilon gene products critical to embryonic and eye development (Philip, 2001).

The distinct expression of leo transcripts in adult ellipsoid body and thorax indicates that LeoI and LeoII may have isoform-specific functions in these tissues and suggest that structural differences between the two isoforms may be reflected in functional specificity. The two Leo isoforms differ by five amino acids in the variable sixth alpha helix. The first two unique amino acids in LeoII (K, N in place of Q, T) are never found at that position among metazoan 14-3-3 isotypes. The third substitution (E in place of D) is present in the vertebrate zeta, beta, tau, eta and sigma isotypes and the two Caenorhabditis elegans isotypes. Finally, the last two amino acids (A, T in place of S, G) are present in both yeast isotypes but not among animal 14-3-3s. Thus, the LeoII isoform appears to be a unique zeta isotype. Helix 6 does not appear to be involved in phosphopeptide binding or dimerization. It is unclear then whether the differences between LeoI and LeoII result in differential dimerization or substrate engagement. The mushroom bodies apparently contain both LeoI and LeoII isoforms and the ellipsoid body contains only LeoI. However, both isoforms rescue equally the olfactory learning and memory deficits of leo mutants; thus, they do not appear to have isoform-specific functions pertinent to these processes. Alternatively, subtle functional differences may have been concealed by elevated transgene expression and the accumulation of a single isoform in the mushroom bodies (Philip, 2001).

Collectively, the data indicate that monomers and homodimers of either Leo isoform and/or heterodimers with D14-3-3epsilon are capable of similar physiological roles essential for learning and memory. The results demonstrate that Leo proteins do not contribute to postembryonic developmental processes in the brain. This is expected to enable investigation and identification of signaling molecules engaged by each isoform in the adult mushroom body and ellipsoid body, which in turn may reveal functional differences among them. The role of Raf1 and the Ras/Raf cascade, which requires leonardo function for signaling in developmental processes, is of particular interest. Furthermore, these results establish an acute role for 14-3-3 proteins in behavioral neuroplasticity, and, by virtue of the high degree of conservation and similarly elevated neuronal expression, are directly applicable to 14-3-3 function in vertebrates (Philip, 2001).

Cell cycle roles for two 14-3-3 proteins during Drosophila development

Drosophila 14-3-3γ and 14-3-3ζ proteins have been shown to function in RAS/MAP kinase pathways that influence the differentiation of the adult eye and the embryo. Because 14-3-3 proteins have a conserved involvement in cell cycle checkpoints in other systems, it was asked (1) whether Drosophila 14-3-3 proteins also function in cell cycle regulation, and (2) whether cell proliferation during Drosophila development has different requirements for the two 14-3-3 proteins. Antibody staining for 14-3-3 family members is cytoplasmic in interphase and perichromosomal in mitosis. Using mutants of cyclins, Cdk1 and Cdc25string to manipulate Cdk1 activity, it was found that the localization of 14-3-3 proteins is coupled to Cdk1 activity and cell cycle stage. Relocalization of 14-3-3 proteins with cell cycle progression suggested cell-cycle-specific roles. This notion is confirmed by the phenotypes of 14-3-3γ and 14-3-3ζ mutants: 14-3-3γ is required to time mitosis in undisturbed post-blastoderm cell cycles and to delay mitosis following irradiation; 14-3-3ζ is required for normal chromosome separation during syncytial mitoses. A model is suggested in which 14-3-3 proteins act in the undisturbed cell cycle to set a threshold for entry into mitosis by suppressing Cdk1 activity, to block mitosis following radiation damage and to facilitate proper exit from mitosis (Su, 2001).

In a previous study of 14-3-3γ localization in the embryo, this protein was reported to become nuclear-localized in infolding cells (Tien, 1999). However, a close examination of the published data revealed that the localization was in pre-mitotic cells (the publication featured mitotic domain 14 that borders the ventral furrow). In fact, a close correspondence of cells that show nuclear-localized 14-3-3γ in this publication (Tien, 1999) and cells that compose the mitotic domains is what led to further examination of the role of 14-3-3 proteins in the cell cycle. Using the same antibody and the same conditions, similar staining patterns were demonstrated (Tien, 1999). A different interpretation of these data is being offered. No correlation of the localized staining with the movement of cells or folding of the epithelium was found. Instead, the findings that 14-3-3 proteins localize to the perichromosomal region during mitosis and that this localization is coupled to Cdk1 activity demonstrate that localization is coupled to cell cycle progression and suggest that 14-3-3 proteins have a cell cycle role (Su, 2001).

One striking set of data presented in this study concern the localization of 14-3-3 proteins to the neighborhood of chromosomes in mitosis. Although the perinuclear localization of Drosophila 14-3-3 proteins is unprecedented, the interphase location and activity are consistent with reports from other systems. S. pombe Rad24 remains exclusively cytoplasmic throughout the cell cycle and this localization appears to be important for blocking mitosis upon checkpoint activation. Similarly, it has been proposed that cytoplasmic human 14-3-3sigma inhibits mitosis by retaining Cdk1/cyclin B in the cytoplasm (Chan, 1999). Like their homologs in other systems, Drosophila 14-3-3 proteins are cytoplasmic in interphase, and analysis of mutations suggests that Drosophila 14-3-3γ also inhibits entry into mitosis in response to activation of DNA damage checkpoint in embryos. This is in agreement with its proposed role in other species and consistent with a recent report (Brodsky, 2000) of a role for 14-3-3γ in preventing mitosis after DNA damage in Drosophila larvae (Su, 2001).

In addition, observations indicate a role for 14-3-3γ in the normal timing of embryonic mitoses. The precise schedule of mitotic times of cells in various positions in the Drosophila embryo made possible detection of deviations from normal timing that are as small as a few minutes. Defects can occur in the normally rigid stereotypical order with which different regions of the embryo progress into mitosis. For example, recent reports described the premature mitosis of mesodermal cells, normally domain 10, in a mutant tribbles. When embryos deficient in 14-3-3γ were examined, a different type of timing defect was found. The normal order of the mitotic domains was retained, but the entire schedule of mitosis was advanced relative to germ-band extension, a major morphological marker of developmental progression. Because there was no detectable slowing of germ-band extension in 14-3-3γ mutant embryos, it is infered that mitosis is advanced in embryos that lack 14-3-3γ. Thus, 14-3-3γ might set physiologically relevant thresholds for entry into mitosis in Drosophila, and this activity might be amplified in response to irradiation. S. pombe mutants in a 14-3-3 homolog show smaller cell size at division; because cellular growth in this organism occurs mainly in G2, it has been proposed that G2 is shorter in these 14-3-3 mutants (Ford, 1994), although precise measurements of this period have not been reported. Thus, it remains to be seen whether 14-3-3 proteins have a similar ability to set the threshold for normal mitosis in other species where only its checkpoint function has been detected (Su, 2001).

14-3-3ζ mutants show defective mitoses in the syncytium, indicating a requirement for this protein in syncytial divisions. Embryos that lack checkpoint functions such as Grapes (Chk1 homolog) and Mei-41 (an ATR homolog) also show mitotic defects, and it has been proposed that these defects are secondary to entry into mitosis with unreplicated DNA. However, loss of 14-3-3ζ functions affects early cycles. By contrast, the dramatic phenotypes of checkpoint defects occur at later syncytial stages (around cycle 12) when checkpoints are thought to become essential to postpone mitosis as S phase takes longer to complete. Thus, the early phenotype of 14-3-3ζ mutant embryos suggests that 14-3-3ζ has roles beyond its likely function in the checkpoint. Perhaps, like 14-3-3γ, 14-3-3ζ might contribute to the normal timing of mitosis even when checkpoints are not operating. Alternatively, incomplete separation of chromosomes in 14-3-3ζ mutants could indicate a more direct involvement of 14-3-3ζ in mitotic progression, an idea that is supported by the localization of the proteins around the mitotic chromosomes and their dispersal after chromosome separation. A direct test of these models will require specific inactivation of 14-3-3ζ in mitosis (as opposed to interphase) (Su, 2001).

Drosophila 14-3-3γ and 14-3-3ζ have documented roles in RAS signaling. Recent data implicate a MAP kinase pathway in cell cycle control in Xenopus, raising the possibility that Drosophila 14-3-3 proteins function through a MAPK pathway to affect their cell cycle roles. This is thought to be unlikely because treatment of Drosophila embryos with pharmacological inhibitors of MAPK pathway did not phenocopy either 14-3-3γ or 14-3-3ζ mutations (Su, 2001).

Regardless of the mechanism of action of 14-3-3ζ, it is notable that it has essential cell cycle roles in the absence of perturbations that normally provoke checkpoint responses. This reinforces other findings in Drosophila and in mammals that suggest that functions normally considered to be checkpoint functions have essential roles in regulating the cell cycle early in development (Su, 2001).

Based on the cytoplasmic localization of 14-3-3γ and cyclin/Cdk1 during interphase, it is proposed that 14-3-3γ acts to keep Cdk1 in check during interphase. As Cdk1 becomes active (owing to the accumulation of its activator Stg or after recovery from DNA damage) and cells enter mitosis, accumulating cyclin/Cdk1 activity promotes and maintains, probably indirectly, 14-3-3 protein localization near chromosomes. Upon the transition to anaphase, the localized 14-3-3 proteins can contribute to chromosome separation. The decline in Cdk1 activity allows 14-3-3 proteins to return to their interphase distribution. Thus, during interphase, 14-3-3γ can act to keep Cdk1 inactive in the cytoplasm but, once Cdk1 is active, it can act in turn to localize 14-3-3 proteins in preparation for their action during the exit from mitosis. No physical interaction has been detected between 14-3-3 proteins and Drosophila homologs of cell cycle regulators known to interact with 14-3-3 proteins in other systems (Cdc25string and cyclin B). Thus, understanding the mechanism of 14-3-3 action might require the identification of novel target molecules (Su, 2001).

The results do not rule out the possibility that 14-3-3ζ also functions to regulate the entry into mitosis in cellular embryos. This possibility cannot be addressed because 14-3-3ζ mutants arrest before G2/M control is first seen in embryogenesis, and the fraction of embryos that do progress to cellular stages are too defective with respect to cell cycle progression and gastrulation. In addition, the fact that these embryos progressed to cellular stages might reflect an incomplete loss of maternal 14-3-3ζ, thus precluding meaningful experiments. What is certain, however, is that 14-3-3γ cannot substitute for 14-3-3ζ during the nuclear divisions of syncytial stages, and that 14-3-3ζ cannot substitute 14-3-3γ for regulating the entry into mitosis during cellular stages (Su, 2001).

In summary, three lines of data indicate that Drosophila 14-3-3 proteins function in normal cell cycle progression, in addition to checkpoint regulation. These are: (1) cell cycle stage specific localization, which is dictated by Cdk1; (2) advancement of mitotic entry in 14-3-3γ mutants; and (3) defective mitoses in 14-3-3ζ mutants. This is the first clear evidence for the requirement for 14-3-3 proteins in normal mitosis in a eukaryote. Furthermore, the fact that mutations in two 14-3-3 proteins lead to different outcomes and at different stages in embryogenesis indicates that these proteins are not functionally redundant. Instead, the results provide strong evidence that, during metazoan development, cell division and its regulation might have different requirements for two members of the 14-3-3 family (Su, 2001).


REFERENCES

Acevedo, S. F., Tsigkari, K. K., Grammenoudi, S. and Skoulakis, E. M. (2007). In vivo functional specificity and homeostasis of Drosophila 14-3-3 proteins. Genetics 177(1): 239-53. PubMed Citation: 17660572

Al-Hakim, A. K., et al. (2005). 14-3-3 cooperates with LKB1 to regulate the activity and localization of QSK and SIK. J. Cell Sci. 118(Pt 23): 5661-73. 16306228

Basu, S., et al. (2003). Akt phosphorylates the Yes-associated protein, YAP, to induce interaction with 14-3-3 and attenuation of p73-mediated apoptosis. Mol. Cell 11: 11-23. 12535517

Benton, R., Palacios, I. M. and St Johnston, D. (2002). Drosophila 14-3-3/PAR-5 is an essential mediator of PAR-1 function in axis formation. Developmental Cell 3: 659-671. 12431373

Benton, R. and Johnston, D. S. (2003). Drosophila PAR-1 and 14-3-3 inhibit Bazooka/PAR-3 to establish complementary cortical domains in polarized cells. Cell 115: 691-704. PubMed Citation: 14675534

Benzing, T., Kottgen, M., Johnson, M., Schermer, B., Zentgraf, H., Walz, G. and Kim, E. (2002). Interaction of 14-3-3 protein with regulator of G protein signaling 7 is dynamically regulated by tumor necrosis factor-alpha. J. Biol. Chem. 277: 32954-32962. 12077120

Bertram, P. G., et al. (1998). The 14-3-3 proteins positively regulate rapamycin-sensitive signaling. Curr. Biol. 8(23): 1259-1267. PubMed Citation: 9822578

Broadie, K., et al. (1997). Leonardo, a Drosophila 14-3-3 protein involved in learning, regulates presynaptic function. Neuron 19(2): 391-402. PubMed Citation: 9292728

Bunney, T. D., De Boer, A. H. and Levin, M. (2003). Fusicoccin signaling reveals 14-3-3 protein function as a novel step in left-right patterning during amphibian embryogenesis. Development 130: 4847-4858. 12930777

Cacace, A. M., et al. (1999). Identification of constitutive and ras-inducible phosphorylation sites of KSR: implications for 14-3-3 binding, mitogen-activated protein kinase binding, and KSR overexpression. Mol. Cell. Biol. 19(1): 229-40. PubMed Citation: 9858547

Chan, H. C., Wu, W. L., So, S. C., Chung, Y. W., Tsang, L. L., Wang, X. F., Yan, Y. C., Luk, S. C., Siu, S. S., Tsui, S. K. et al. (2000). Modulation of the Ca2+-activated Cl- channel by 14-3-3epsilon. Biochem. Biophys. Res. Commun. 270: 581-587. 10753667

Chang, H. C. and Rubin, G. M. (1997). 14-3-3 epsilon positively regulates Ras-mediated signaling in Drosophila. Genes Dev. 11: 1132-1139. 9159394 >

Chen, L., Liu, T. H. and Walworth, N. C. (1999). Association of chk1 with 14-3-3 proteins is stimulated by DNA damage. Genes Dev. 13(6): 675-85. PubMed ID: 10090724

Clark, G. J., et al. (1997). 14-3-3 zeta negatively regulates raf-1 activity by interactions with the Raf-1 cysteine-rich domain. J. Biol. Chem. 272(34): 20990-20993. PubMed ID: 9261098

Craparo, A., Freund, R. and Gustafson, T. A. (1997). 14-3-3 (epsilon) interacts with the insulin-like growth factor I receptor and insulin receptor substrate I in a phosphoserine-dependent manner. J Biol Chem 272 (17): 11663-11669. PubMed ID: 9111084

Cuenca, A. A., et al. (2003). Polarization of the C. elegans zygote proceeds via distinct establishment and maintenance phases. Development 130: 1255-1265. 12588843

Datta, S. R., et al. (1997). Akt phosphorylation of BAD couples survival signals to the cell-intrinsic death machinery. Cell 91: 231-241

DeYoung, M. P., et al. (2008). Hypoxia regulates TSC1/2-mTOR signaling and tumor suppression through REDD1-mediated 14-3-3 shuttling. Genes Dev. 22: 239-251. PubMed Citation: 18198340

Forbes, K. C., Humphrey, T. and Enoch, T. (1998). Suppressors of cdc25p overexpression identify two pathways that influence the G2/M checkpoint in fission yeast. Genetics 150(4): 1361-75.

Geiger, J. C., Lipka, J., Segura, I., Hoyer, S., Schlager, M. A., Wulf, P. S., Weinges, S., Demmers, J., Hoogenraad, C. C. and Acker-Palmer, A. (2014). The GRIP1/14-3-3 pathway coordinates cargo trafficking and dendrite development. Dev Cell 28: 381-393. PubMed ID: 24576423

Honda, R., Ohba, Y. and Yasuda, H. (1997). 14-3-3 zeta protein binds to the carboxyl half of mouse wee1 kinase. Biochem Biophys Res Commun 230 (2): 262-265

Hurd, T. W., et al. (2003). Phosphorylation-dependent binding of 14-3-3 to the polarity protein Par3 regulates cell polarity in mammalian epithelia. Curr. Biol. 13: 2082-2090. 14653998

Jin, J., et al. (2004). Proteomic, functional, and domain-based analysis of in vivo 14-3-3 binding proteins involved in cytoskeletal regulation and cellular organization. Curr. Biol. 14: 1436-1450. 15324660

Kanai, F., et al. (2000). Taz: a novel transcriptional co-activator regulated by interactions with 14-3-3 and pdz domain proteins. EMBO J. 19: 6778-6791. 11118213

Karam, C. S., Kellner, W. A., Takenaka, N., Clemmons, A. W. and Corces, V. G. (2010). 14-3-3 mediates histone cross-talk during transcription elongation in Drosophila. PLoS Genet. 6: e1000975. Pubmed: 20532201

Kellner, W. A., Ramos, E., Van Bortle, K., Takenaka, N. and Corces, V. G. (2012). Genome-wide phosphoacetylation of histone H3 at Drosophila enhancers and promoters. Genome Res. 22: 1081-1088. Pubmed: 22508764

Kim, H. H., et al. (2008). Nuclear HuR accumulation through phosphorylation by Cdk1 Genes Dev. 22: 1804-1815. PubMed Citation: 18593881

Kockel, L., et al. (1997). Requirement for Drosophila 14-3-3 eta in Raf-dependent photoreceptor development. Genes Dev. 11: 1140-47

Koga, Y. and Ikebe, M. (2007). A novel regulatory mechanism of myosin light chain phosphorylation via binding of 14-3-3 to myosin phosphatase. Mol Biol Cell. 19(3): 1062-71. Medline abstract: 18094049

Kumagai, A., Yakowec, P. S. and Dunphy, W. G. (1998). 14-3-3 proteins act as negative regulators of the mitotic inducer cdc25 in Xenopus egg extracts. Mol. Biol. Cell 9(2): 345-354

Lee, J., Kumagai, A. and Dunphy, W. G. (2001). Positive regulation of Wee1 by Chk1 and 14-3-3 proteins. Mol. Biol. Cell 12: 551-563. 11251070

Li, W., et al. (1997). The Drosophila 14-3-3 protein Leonardo enhances Torso signaling through D-Raf in a Ras 1-dependent manner. Development 124(20): 4163-4171

Ling, C., Zuo, D., Xue, B., Muthuswamy, S. and Muller, W. J. (2010). A novel role for 14-3-3sigma in regulating epithelial cell polarity. Genes Dev. 24(9): 947-56. PubMed Citation: 20439433

Liu, Y. C., et al. (1996). Activation-modulated association of 14-3-3 proteins with Cbl in T cells. J Biol Chem 271 (24): 14591-14595

Liu, Y. C., et al. (1997). Serine phosphorylation of Cbl induced by phorbol ester enhances its association with 14-3-3 proteins in T cells via a novel serine-rich 14-3-3-binding motif. J Biol Chem 272 (15): 9979-9985

Lo, M.-C., et al. (2004). Phosphorylation by the ß-Catenin/MAPK complex promotes 14-3-3-mediated nuclear export of TCF/POP-1 in signal-responsive cells in C. elegans. Cell 117: 95-106. 15066285

Lopez-Girona, A., et al. (1999). Nuclear localization of Cdc25 is regulated by DNA damage and a 14-3-3 protein. Nature 397(6715): 172-5. PubMed ID: 9923681

Lu, M. S. and Prehoda, K. E. (2013). A NudE/14-3-3 pathway coordinates dynein and the kinesin Khc73 to position the mitotic spindle. Dev Cell 26: 369-380. PubMed ID: 23987511

Margolis, S. S., et al. (2006). Role for the PP2A/B56delta phosphatase in regulating 14-3-3 release from Cdc25 to control mitosis. Cell 127(4): 759-73. Medline abstract: 17110335

McCaffrey, L. M. and Macara, I. G. (2009). The Par3/aPKC interaction is essential for end bud remodeling and progenitor differentiation during mammary gland morphogenesis. Genes Dev. 23: 1450-1460. PubMed Citation: 19528321

Meller, N., et al. (1996). Direct interaction between protein kinase C theta (PKC theta) and 14-3-3 tau in T cells: 14-3-3 overexpression results in inhibition of PKC theta translocation and function. Mol Cell Biol 16 (10): 5782-5791

Morton, D. G., et al. (2002). The Caenorhabditis elegans par-5 gene encodes a 14-3-3 protein required for cellular asymmetry in the early embryo. Dev. Bio. 241: 47-58. 11784094

Murakami, K., Situ, S. Y. and Eshete, F. (1996). A gene variation of 14-3-3 zeta isoform in rat hippocampus. Gene 179 (2): 245-249

Muslin, A. J., et al. (1996). Interaction of 14-3-3 with signaling proteins is mediated by the recognition of phosphoserine. Cell 84 (6): 889-897

Nagata-Ohashi, K., et al. (2004). A pathway of neuregulin-induced activation of cofilin-phosphatase Slingshot and cofilin in lamellipodia. J. Cell Biol. 165(4): 465-71. 15159416

Nutt, L. K., et al. (2005). Metabolic regulation of oocyte cell death through the CaMKII-mediated phosphorylation of caspase-2. Cell 123: 89-103. PubMed Citation: 16213215

Nutt, L. K., et al. (2010). Metabolic control of oocyte apoptosis mediated by 14-3-3zeta-regulated dephosphorylation of caspase-2. Dev. Cell 16(6): 856-66. PubMed Citation: 19531356

Obsil, T., et al. (2001). Crystal structure of the 14-3-3zeta:Serotonin N-acetyltransferase complex: A role for scaffolding in enzyme regulation. Cell 105: 257-267. 11336675

Ogihara, T., et al. (1997). 14-3-3 protein binds to insulin receptor substrate-1, one of the binding sites of which is in the phosphotyrosine binding domain. J. Biol. Chem. 272(40): 25267-25274. PubMed Citation: 9312143

Ory, S., et al. (2003). Protein phosphatase 2A positively regulates ras signaling by dephosphorylating KSR1 and Raf-1 on critical 14-3-3 binding sites. Curr. Biol. 13: 1356-1364. 12932319

Peng, C.Y., et al. (1997). Mitotic and G2 checkpoint control: Regulation of 14-3-3 protein binding by phosphorylation of Cdc25C on serine-216. Science 277: 1501-1505. PubMed Citation: 9278512

Pflieger, D., et al. (2008). Quantitative proteomic analysis of protein complexes: concurrent identification of interactors and their state of phosphorylation. Mol. Cell Proteomics. 7(2): 326-46. PubMed Citation: 17956857

Philip, N., Acevedo, S. F. and Skoulakis, E. M. C. (2001). Conditional rescue of olfactory learning and memory defects in mutants of the 14-3-3zeta gene leonardo. J. Neurosci. 21(21): 8417-8425. 11606630

Regnard, C., (2011). Global analysis of the relationship between JIL-1 kinase and transcription. PLoS Genet 7: e1001327. Pubmed: 21423663

Rittinger, K., et al. (1999). Structural analysis of 14-3-3 phosphopeptide complexes identifies a dual role for the nuclear export signal of 14-3-3 in ligand binding. Mol. Cell 4: 153-166. PubMed Citation: 10488331

Roberts, R. L., Mosch, H. U. and Fink, G. R. (1997). 14-3-3 proteins are essential for RAS/MAPK cascade signaling during pseudohyphal development in S. cerevisiae. Cell 89(7): 1055-1065

Rommel, C., et al. (1996). Activated Ras displaces 14-3-3 protein from the amino terminus of c-Raf-1. Oncogene 12 (3): 609-619

Roy, S., et al. (1998). 14-3-3 facilitates Ras-dependent Raf-1 activation in vitro and in vivo. Mol. Cell. Biol. 18(7): 3947-3955

Santoro, M. M., Gaudino, G. and Marchisio, P. C. (2003). The MSP receptor regulates alpha6beta4 and alpha3beta1 integrins via 14-3-3 proteins in keratinocyte migration. Dev. Cell 5: 257-271. 12919677

Sato, M., et al. (2002). 14-3-3 Protein interferes with the binding of RNA to the phosphorylated form of fission yeast meiotic regulator Mei2p. Curr. Biol. 12: 141-145. 11818066

Skoulakis, E. M. C. and Davis, R. L. (1996). Olfactory learning deficits in mutants for leonardo, a Drosophila gene encoding a 14-3-3 protein. Neuron 17: 931-944. PubMed Citation: 8938125

Su, T. T., et al. (2001). Cell cycle roles for the two 14-3-3 proteins during Drosophila development. J. Cell Sci. 114: 3445-3462. PubMed Citation: 11682604

Suzuki, A., et al. (2004). aPKC acts upstream of PAR-1b in both the establishment and maintenance of mammalian epithelial polarity. Curr. Biol. 14: 1425-1435. 15324659

Swanson, K. D. and Ganguly, R. (1992). Characterization of a Drosophila melanogaster gene similar to the mammalian genes encoding the tyrosine/tryptophan hydroxylase activator and protein kinase C inhibitor proteins. Gene 113 (2): 183-190

Thorson, J. A., et al. (1998). 14-3-3 proteins are required for maintenance of Raf-1 phosphorylation and kinase activity. Mol. Cell. Biol. 18(9): 5229-38

Tien, A. C., Hsei, H. Y. and Chien, C. T. (1999). Dynamic expression and cellular localization of the Drosophila 14-3-3epsilon during embryonic development. Mech. Dev. 81(1-2): 209-12

Tokumitsu, H., et al. (2005). Phosphorylation of Numb family proteins. Possible involvement of Ca2+/calmodulin-dependent protein kinases. J. Biol. Chem. 280(42): 35108-18. 16105844

Traweger, A., Wiggin, G., Taylor, L., Tate, S. A., Metalnikov, P. and Pawson, T. (2008). Protein phosphatase 1 regulates the phosphorylation state of the polarity scaffold Par-3. Proc. Natl. Acad. Sci. 105(30): 10402-7. PubMed Citation: 18641122

Tzivion, G., Luo, Z. and Avruch, J. (1998). A dimeric 14-3-3 protein is an essential cofactor for Raf kinase activity. Nature 394(6688): 88-92

Vincenz, C. and Dixit, V. M. (1996). 14-3-3 proteins associate with A20 in an isoform-specific manner and function both as chaperone and adapter molecules. J Biol Chem 271 (33): 20029-20034.

Wakui, H., et al. (1997). Interaction of the ligand-activated glucocorticoid receptor with the 14-3-3 eta protein. J Biol Chem 272 (13): 8153-8156

Wang, H. G., et al. (1996). Bcl-2 interacting protein, BAG-1, binds to and activates the kinase Raf-1. Proc. Natl. Acad. Sci. 93 (14): 7063-7068

Wang, W. and Shakes, D. C. (1994). Isolation and sequence analysis of a Caenorhabditis elegans cDNA which encodes a 14-3-3 homologue. Gene 147 (2): 215-218

Wang, W. and Shakes, D. C. (1996). Molecular evolution of the 14-3-3 protein family. J. Mol. Evol. 43 (4): 384-398

Xin, Y., Lu, Q. and Li, Q. (2010). 14-3-3sigma controls corneal epithelial cell proliferation and differentiation through the Notch signaling pathway. Biochem. Biophys. Res. Commun. 392: 593-598. PubMed Citation: 20100467

Xing, H., Kornfeld, K., and Muslin, A. J. (1997). The protein kinase KSR interacts with 14-3-3 protein and Raf. Curr. Biol. 7 (5): 294-300

Xing, H., et al. (2000). 14-3-3 proteins block apoptosis and differentially regulate MAPK cascades. EMBO J. 19: 349-358.

Yang, J., et al. (1999). Maintenance of G2 arrest in the Xenopus oocyte: a role for 14-3-3-mediated inhibition of Cdc25 nuclear import. EMBO J. 18(8): 2174-2183

Yano, S., Tokumitsu, H., Soderling, T. R. (1998). Calcium promotes cell survival through CaM-K kinase activation of the protein-kinase-B pathway. Nature 396(6711): 584-7

Yuan, H., Michelsen, K. and Schwappach, B. (2003). 14-3-3 dimers probe the assembly status of multimeric membrane proteins. Curr. Biol. 13: 638-646. 12699619

Zeng, Y., et al. (1998). Replication checkpoint requires phosphorylation of the phosphatase Cdc25 by Cds1 or Chk1. Nature 395(6701): 507-10

Zha, J., et al. (1996). Serine phosphorylation of death agonist BAD in response to survival factor results in binding to 14-3-3 not BCL-X(L) Cell 87 (4): 619-628

Zhang, L., et al. (1997). Raf-1 kinase and exoenzyme S interact with 14-3-3zeta through a common site involving lysine 49. J Biol Chem 272 (21): 13717-13724

Zhang, S. H., et al. (1997). Serine phosphorylation-dependent association of the band 4.1-related protein-tyrosine phosphatase PTPH1 with 14-3-3beta protein. J. Biol. Chem. 272(43): 27281-27287

Zhang, L., Chen, J. and Fu, H. (1999). Suppression of apoptosis signal-regulating kinase 1-induced cell death by 14-3-3 proteins. Proc. Natl. Acad. Sci. 96(15): 8511-5

Zhou, Q., et al. (2010). 14-3-3 coordinates microtubules, Rac, and myosin II to control cell mechanics and cytokinesis. Curr Biol. 20(21): 1881-9. PubMed Citation: 20951045

Zhou, Y., et al. (1999). A dynamically regulated 14-3-3, Slob, and Slowpoke potassium channel complex in Drosophila presynaptic nerve terminals. Neuron 22(4): 809-18


14-3-3zeta/leonardo: Biological Overview | Evolutionary Homologs | Regulation | Developmental Biology | Effects of Mutation

date revised: 20 April 2014

 

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

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