Skip to main content

Nucleo-cytoplasmic shuttling of murine RBPJ by Hairless protein matches that of Su(H) protein in the model system Drosophila melanogaster

Abstract

CSL transcription factors are central to signal transduction in the highly conserved Notch signaling pathway. CSL acts as a molecular switch: depending on the cofactors recruited, CSL induces either activation or repression of Notch target genes. Unexpectedly, CSL depends on its cofactors for nuclear entry, despite its role as gene regulator. In Drosophila, the CSL homologue Suppressor of Hairless (Su(H)), recruits Hairless (H) for repressor complex assembly, and eventually for nuclear import. We recently found that Su(H) is subjected to a dynamic nucleo-cytoplasmic shuttling, thereby strictly following H subcellular distribution. Hence, regulation of nuclear availability of Su(H) by H may represent a new layer of control of Notch signaling activity. Here we extended this work on the murine CSL homologue RBPJ. Using a ‘murinized’ fly model bearing RBPJwt in place of Su(H) at the endogenous locus we demonstrate that RBPJ protein likewise follows H subcellular distribution. For example, overexpression of a H*NLS3 protein variant defective of nuclear import resulted in a cytosolic localization of RBPJ protein, whereas the overexpression of a H*NES protein variant defective in the nuclear export signal caused the accumulation of RBPJ protein in the nucleus. Evidently, RBPJ is exported from the nucleus as well. Overall these data demonstrate that in our fly model, RBPJ is subjected to H-mediated nucleo-cytoplasmic shuttling as is Su(H). These data raise the possibility that nuclear availability of mammalian CSL proteins is likewise restricted by cofactors, and may hence present a more general mode of regulating Notch signaling activity.

Graphical abstract

Background

Development as well as tissue homeostasis of higher eumetazoa depends on inter-cellular communication mediated by the Notch signaling pathway. Accordingly, the Notch signaling pathway is highly conserved in the evolution of invertebrates and vertebrates alike [1,2,3]. Upon binding of one of its ligands, the Notch receptor undergoes cleavage releasing the Notch intracellular domain NICD. Together with several cofactors, NICD assembles a transcriptional activator complex switching gene expression, and eventually cell fate, in the signal-receiving cell [3,4,5,6,7]. Pivotal to Notch target gene regulation is the DNA binding protein CSL; CSL is an acronym for mammalian CBF1/ RBPJ, for Drosophila Su(H) and for Caenorhabditis Lag1. Crystal structure analyses of the trimeric activator complex revealed that NICD contacts CSL with its RAM-domain and ankyrin repeats, allowing recruitment of the coactivator Mam [8]. In the absence of signal, CSL engages in Notch target gene inhibition by forming a repressor complex on Notch target gene promoters [9, 10]. Several corepressors have been identified in mammals, which compete with NICD for the RAM-binding site within the beta-trefoil domain of CSL [9,10,11,12]. The major antagonist of the Notch signaling pathway in Drosophila is named Hairless (H) [13]. Contrary to most of the mammalian CSL corepressors, H contacts the C-terminal domain of the fly CSL homologue named Suppressor of Hairless (Su(H)) [14,15,16]. By recruiting two general corepressors, Groucho and C-terminal binding protein, the Su(H)-H repressor complex eventually silences Notch target genes [13, 17,18,19]. With SHARP (also named MINT), a functional homologue of H has been identified in vertebrates [9, 11, 20, 21]. SHARP binds CSL in a bipartite manner, i.e. both within the beta-trefoil domain and the C-terminal domain, resembling the interaction of mammalian corepressors as well as of H with CSL [22].

Unexpected for a transcription factor, CSL apparently relies on its cofactors for nuclear entry. For example, mutations of CBF1/RBPJ in the beta-trefoil domain affecting both, the binding of NICD as well as of corepressors, prevented nuclear entry and Notch target gene activation [23]. Similarly, Su(H) nuclear entry depended on NICD in a Drosophila cell culture system; hence it may not enter the nuclear compartment on its own [24, 25]. Moreover, tissue-specific overexpression of H protein caused Su(H) nuclear accumulation, whereas Su(H) protein levels appeared reduced in the absence of H protein [26,27,28]. In fact, it was demonstrated that Su(H) protein stability depends on formation of transcription complexes together with H and NICD, respectively [28].

Subcellular localization of Hairless and suppressor of Hairless protein

We recently addressed the subcellular localization of H and Su(H) proteins in Drosophila tissue and showed that Su(H) protein strictly follows the subcellular localization of H [29]. H protein contains three potential nuclear localization signals NLS1–3, with NLS3 being the most effectual. Accordingly, H*NLS3 mutant protein defective in NLS3 accumulated within the cytosol. In addition, a nuclear export signal NES, juxtaposed to NLS3, proved relevant for the export of H protein from the nuclear compartment. Mutation of the NES resulted in nuclear retention of H*NES protein in larval tissues. Endogenous Su(H) protein co-localized with H protein, i.e. it was cytosolic when the H*NLS3 mutant was overexpressed and nuclear in cells expressing H*NES [29]. A double mutant H*NLS3*NES had an intermediate effect, and either protein distribution resembled the wild type situation, demonstrating the importance of the NES in H and Su(H) export. Overall our data implied, that H mediated shuttling of Su(H) between the nucleo-cytosolic compartments provided a means of regulating Notch activity by restricting nuclear availability of Su(H). Here we asked, whether mammalian CSL protein might be subjected to a similar mode of regulation. The fact that nuclear import of CBF1/RBPJ is dependent on its cofactors as well makes this hypothesis very likely. Moreover, in yeast two-hybrid assays murine CBF1/RBPJ was shown capable of binding H with its C-terminal domain similar to Su(H) [14, 15, 30].

Murine RBPJ protein follows the subcellular distribution of Hairless protein in the fly

To address the potential role for H on nucleo-cytoplasmic shuttling of mammalian CSL, we made use of a ‘murinized’ fly model which we recently established [30]. In these flies, the endogenous Su(H) locus has been replaced by the murine CSL orthologue RBPJ using genome engineering. Interestingly, RBPJwt flies are viable with subtle phenotypes, demonstrating that the murine CSL orthologue can replace the majority of Su(H) activities during fly development [30]. The RBPJwt fly model allowed us to test, whether RBPJ protein is subjected to H-mediated nuclear localization like its fly homologue Su(H), i.e. nuclear import – as expected by a likewise nuclear import of CBF1/RBPJ by corepressors – as well as nuclear export, as uncovered for Su(H) in Drosophila. We applied the Gal4-UAS system [31] for a tissue specific overexpression of H* variants mutant in a nuclear translocation signal, as this setting allows following the distribution of endogenous CSL protein within larval tissue [29]. For the overexpression, we used sd-Gal4 [32] driving UAS-H* transgene expression in the larval salivary glands, where subcellular protein localization can be easily visualized in the cytoplasm and nuclei of the giant cells [29]. To this end, we first combined the RBPJwt bearing 2nd chromosome with the sd-Gal4 line and the UAS-H* transgenes, respectively, to generate driver and effector lines in the RBPJwt genetic background (Fig. 1). Four RBPJwt-bearing effector lines were established: UAS-Hcwt as control, UAS-H*NLS3 defective for nuclear import, UAS-H*NES defective for nuclear export, and UAS-H*NLS3*NES affecting both, import- and export signal [29].

Fig. 1
figure 1

Crossing scheme. Crossing scheme for establishing (a) the driver line sd-Gal4; RBPJwt / CyO-GFP and (b) the effector lines RBPJwt / CyO-GFP; UAS-H* (representing the four different H alleles, UAS-Hcwt, UAS-H*NLS3, UAS-H*NES and UAS-H*NLS3*NES, respectively). Direction of the cross is indicated with males (m), and virgin females (f). Note that sd-Gal4 is X-linked. Crosses of driver and effector lines result in the desired offspring, i.e. third instar larvae homozygous for RBPJwt that can be selected for the absence of the GFP marker for subsequent analysis of salivary glands

Each effector line was crossed with the sd-Gal4; RBPJwt driver line to induce the overexpression of the respective H* mutant protein in the salivary glands of RBPJwt larvae. Staining of the salivary glands then revealed the subcellular distribution of H* and RBPJwt protein, respectively (Fig. 2a). As shown earlier [29], H*NLS3 protein was mostly cytoplasmic, whereas all other H* variants were detected in both, cytosolic and nuclear compartment. Notably, H*NES appeared more strongly enriched in the nucleus than Hcwt and H*NLS3*NES protein (Fig. 2a). As predicted from the subcellular localization of Su(H) [29], RBPJwt protein was cytosolic when H*NLS3 protein was overexpressed, and detected in the nuclear compartment as well in the presence of any other H* protein variant (Fig. 2a). To confirm the visual impression, we quantified the staining intensities of confocal micrographs on eight specimen each for every genotype, comprising a minimum of 160 nuclei. Composite Z-stacks crossing the entire gland were segmented into nuclei and cytoplasm, and mean grey values were recorded. The results confirm that murine RBPJwt protein is shuttled by H protein the same way as is Su(H) protein (Fig. 2b). Overexpression of the wild type protein isoform Hcwt caused strong accumulation of RBPJwt protein in the nucleus, and even stronger, when H*NES was overexpressed. In contrast, overexpression of H*NLS3 resulted in the retention of RBPJwt in the cytoplasm, whereas that of H*NLS3*NES allowed RBPJwt protein to re-enter the nucleus (Fig. 2b). Briefly, we observed a nucleo-cytosolic shuttling of RBPJwt protein, which strictly followed H protein distribution in the salivary glands of Drosophila larvae.

Fig. 2
figure 2

Subcellular co-localization of RBPJ and H proteins. a Enlargements of salivary glands derived from homozygous RBPJwt larvae overexpressing the indicated H* protein isoform. Subcellular distribution of H protein is shown in green and of RBPJ protein in magenta; the left panel shows the merge. Size bar represents 50 μm in all panels. The following genotypes are depicted: sd-Gal4/+; RBPJwt / RBPJwt; UAS-Hcwt/+, sd-Gal4/+; RBPJwt / RBPJwt; UAS-H*NLS3/+, sd-Gal4/+; RBPJwt / RBPJwt; UAS-H*NES/+, sd-Gal4/+; RBPJwt / RBPJwt; UAS-H*NLS3*NES/+. b Nuclear to cytoplasmic (n/c) ratio is shown for H protein (green bars) and Su(H) protein (magenta bars), respectively, determined from 8 specimen each indicated as squares. Sample mean and standard deviation is indicated. The dotted line represents equal distribution in both compartments (i.e. nuclear equals cytoplasmic grey value). Hcwt is primarily nuclear, and H*NES even more enriched in nuclei. In contrast, H*NLS3 is located in the cytosol, whereas H*NLS3*NES is detected in the nuclear compartment as well. Note that RBPJwt strictly follows H* subcellular protein distribution. Statistical analysis was performed using ANOVA two-tailed Dunnett’s approach for multiple comparisons relative to the Hcwt control (*p < 0.05; **p < 0.01; ***p < 0.001)

RBPJwt protein accumulated significantly stronger in nuclei upon the overexpression of H*NES (Fig. 2) which is defective in the nuclear export signal [29]. Evidently, RBPJwt is subjected to nuclear export by wild type H protein similar to Su(H). The importance of nuclear export of CSL-H has not been elucidated yet. We know, however, that the H-NES is relevant for fly survival, as in its absence, only a fraction of the animals developed to adulthood [29]. In mouse cells, the tubulin-binding protein RITA induced nuclear export of RBPJ, thereby downregulating Notch-mediated transcription [33]. The more important roles of RITA, however, lie in the regulation of microtubule dynamics during mitosis and cell motility [34, 35]. Albeit its high conservation in the animal kingdom, RITA has no fly homologue. Accordingly, despite binding to Su(H) and tubulin, human RITA has no biological effect on the subcellular distribution or the stability of Su(H) protein in the fly [36]. In contrast to mammalian cells, sequestration of Su(H) by a tubulin-tether in the cytosolic compartment does not occur [36]. Nevertheless, regulation of nuclear availability of CSL proteins appears an important layer of regulation during the transduction of Notch signals in vertebrates and equally in invertebrates.

Conclusion

Nucleo-cytoplasmic shuttling of Su(H) as a means of regulating Notch signaling activity in the fly has been already shown. Here we demonstrate that murine RBPJ is subjected to a likewise dynamic nucleo-cytoplasmic shuttling by H protein in vivo in Drosophila tissue. These data support the hypothesis that nuclear availability of mammalian CSL proteins is restricted by their cofactors, on which they depend for nuclear import. Moreover, murine RBPJ protein is also subjected to nuclear export by H protein. Overall, our data demonstrate the requirement of corepressors for CLS nuclear translocation, emphasizing the additional layer of regulation at the level of nuclear availability.

Methods

The genome engineered fly stock RBPJwt / CyO-GFP contains murine RBPJ cDNA (isoform 1; the N-terminal 128 codons are derived from Su(H) fused at Val-codon 81 to RBPJ) in place of wild type Su(H) [30]. The stock was combined with sd-Gal4 (BL8609) to generate a driver line, and with either UAS-Hcwt, UAS-H*NLS3, UAS-H*NES or UAS-H*NLS3*NES [29] to generate an effector line, by standard genetic crosses as outlined in Fig. 1. To this end, we made use of the dominant markers snaSco (BL9325) and L2 (BL319) [37], and a doubly balanced cycGHR7 allele [38], to be able to follow unambiguously every chromosome through all generations. Driver and effector lines were crossed, and offspring reared at 25 °C to eventually analyse the salivary glands at third instar larval stage. The homozygous RBPJwt animals were recognized by the lack of GFP, otherwise marking the heterozygous siblings due to the CyO-GFP (BL9325) marker. A Leica MZ FLIII UV stereo-microscope (Leica, Wetzlar, Germany) illuminated with CoolLED pE-300white (AHF, Tübingen, Germany) was used for the purpose of selecting the larvae.

Respective UAS-constructs were expressed in the developing salivary glands using sd-Gal4. To visualize H and RBPJ protein expression, immuno-cytochemistry on third instar larval salivary glands was performed as outlined before, with a 20 min fixation with 4% paraformaldehyde [29]. As primary antibodies, we used guinea pig anti-Hairless A (1:500) [27] and rabbit anti-RBPSUH (1:200) (D10A4; Cell Signaling Technology, Cambridge, UK). Goat secondary antibodies (1:250), coupled to FITC or Cy3 were obtained from Jackson Immuno-Research (Dianova, Hamburg, Germany). Fluorescently labelled tissue was mounted in Vectashield (Vector labs, Eching, Germany). Pictures were taken with a Zeiss Axioskop (Carl Zeiss, Jena, Germany), coupled to a BioRad MRC1024 confocal microscope (Carl Zeiss, Jena, Germany; O.S.T.I. microscopy, Milano, Italy) using LaserSharp 2000™ software. The presented figures were created using ImageJ, PhotoPaint and CorelDraw software.

Quantification of H and Su(H) protein in salivary glands overexpressing the specific H* nuclear localization mutant was performed based on confocal micrographs using Image J software. Z-stacks crossing the entire glands with 1 μm increments were used for maximum projection. Composite images were segmented into nuclei and cytoplasm. Nuclei were defined as region of interest, and subtracted from the outline of the whole gland, defining the cytoplasm. Mean grey values of nuclei and corresponding cytoplasm of the entire gland were recorded [29]. Eight glands each with a total of at least 160 nuclei were analyzed. Statistical significance was determined by ANOVA two-tailed Dunnett’s approach for multiple comparisons.

Availability of data and materials

All data generated or analyzed during this study are included in this published article.

Abbreviations

CBF1:

C promoter binding factor 1 (mammalian)

CSL:

CBF1/RBPJ, Su(H), Lag-1 (acronym)

Gal4:

galactose responsive transcription factor (from S. cerevisiae)

GFP:

Green fluorescent protein (from A. Victoria)

H:

Hairless (from D. melanogaster)

HDAC1:

Histone deacetylase 1 (mammalian)

Lag1:

Abnormal cell lineage 12 (lin-12) and abnormal germ line proliferation phenotype-a (glp-1) (from C. elegans)

Mam:

Mastermind from D. melanogaster; corresponds to mammalian MamL (mastermind like transcriptional coactivator)

MINT:

Msx2-interacting nuclear target (mammalian)

NCOR2:

Nuclear receptor corepressor 2 (mammalian)

NES:

Nuclear export signal

NICD:

Notch intracellular domain

NLS:

Nuclear localization signal

RAM:

RBP-Jkappa-associated module (part of NICD)

RBPJ:

Recombination signal binding protein for immunoglobulin kappa J region (mammalian)

RITA:

RBPJ interacting and tubulin associated, also named ZNF331 (mammalian)

SHARP:

SMRT/HDAC1-associated repressor protein (mammalian)

SMRT (or NCOR2):

Silencing mediator for retinoid and thyroid-hormone receptors (mammalian)

Su(H):

Suppressor of Hairless from D. melanogaster

UAS:

Upstream activating sequences (Gal4 target sequence)

ZNF331:

Zinc finger protein 331 (mammalian)

References

  1. Kelleher FC, Fennelly D, Rafferty M. Common critical pathways in embryogenesis and cancer. Acta Oncol. 2006;45(4):375–88. https://doi.org/10.1080/02841860600602946.

    Article  CAS  PubMed  Google Scholar 

  2. Siebel C, Lendahl U. Notch signaling in development, tissue homeostasis, and disease. Physiol Rev. 2017;97(4):1235–94. https://doi.org/10.1152/physrev.00005.2017.

    Article  CAS  PubMed  Google Scholar 

  3. Kopan R, Ilagan MXG. The canonical Notch signaling pathway: unfolding the activation mechanism. Cell. 2009;137(2):216–33. https://doi.org/10.1016/j.cell.2009.03.045.

  4. Bray SJ. Notch signalling in context. Nat Rev Mol Cell Biol. 2016;17(11):722–35. https://doi.org/10.1038/nrm.2016.94.

    Article  CAS  PubMed  Google Scholar 

  5. Bray SJ, Gomez-Lamarca M. Notch after cleavage. Curr Opin Cell Biol. 2018;51:103–9. https://doi.org/10.1016/j.ceb.2017.12.008.

    Article  CAS  PubMed  Google Scholar 

  6. Sjöqvist M, Andersson ER. Do as I say, Not (ch) as I do: lateral control of cell fate. Dev Biol. 2019;447(1):58–70. https://doi.org/10.1016/j.ydbio.2017.09.032.

  7. Gordon WR, Arnett KL, Blacklow SC. The molecular logic of Notch signaling--a structural and biochemical perspective. J Cell Sci. 2008;121(19):3109–19. https://doi.org/10.1242/jcs.035683.

  8. Kovall RA, Blacklow SC. Mechanistic insights into Notch receptor signaling from structural and biochemical studies. Curr Top Dev Biol. 2010;92:31–71. https://doi.org/10.1016/S0070-2153(10)92002-4.

  9. Borggrefe T, Oswald F. The Notch signaling pathway: transcriptional regulation at Notch target genes. Cell Mol Life Sci. 2009;66(10):1631–46. https://doi.org/10.1007/s00018-009-8668-7.

  10. Wang H, Zang C, Liu XS, Aster JC. The role of Notch receptors in transcriptional regulation. J Cell Physiol. 2015;230(5):982–8. https://doi.org/10.1002/jcp.24872.

  11. Maier D. The evolution of transcriptional repressors in the Notch signaling pathway: a computational analysis. Hereditas. 2019;156(1):5. https://doi.org/10.1186/s41065-019-0081-0.

  12. Hall DP, Kovall RA. Structurally conserved binding motifs of transcriptional regulators to Notch nuclear effector CSL. Exp Biol Med (Maywood). 2019;244:1520–9.

    Article  CAS  Google Scholar 

  13. Maier D. Hairless: the ignored antagonist of the Notch signalling pathway. Hereditas. 2006;143(2006):212–21. https://doi.org/10.1111/j.2007.0018-0661.01971.x.

  14. Maier D, Kurth P, Schulz A, Russell A, Yuan Z, Gruber K, Kovall RA, Preiss A. Structural and functional analysis of the repressor complex in the Notch signaling pathway of Drosophila melanogaster. Mol Biol Cell. 2011;22(17):3242–52. https://doi.org/10.1091/mbc.e11-05-0420.

  15. Yuan Z, Praxenthaler H, Tabaja N, Torella R, Preiss A, Maier D, Kovall RA. Structure and function of the Su(H)-Hairless repressor complex, the major antagonist of Notch signaling in Drosophila melanogaster. PLoS Biol. 2016;14(7):e1002509. https://doi.org/10.1371/journal.pbio.1002509.

  16. Borggrefe T, Oswald F. Setting the stage for Notch: the Drosophila Su(H)-Hairless repressor complex. PLoS Biol. 2016;14(7):e1002524. https://doi.org/10.1371/journal.pbio.1002524.

  17. Morel V, Lecourtois M, Massiani O, Maier D, Preiss A, Schweisguth F. Transcriptional repression by Suppressor of Hairless involves the binding of a Hairless-dCtBP complex in Drosophila. Curr Biol. 2001;11(10):789–92. https://doi.org/10.1016/S0960-9822(01)00224-X.

  18. Barolo S, Stone T, Bang AG, Posakony JW. Default repression and Notch signaling: Hairless acts as an adaptor to recruit the corepressors Groucho and dCtBP to Suppressor of Hairless. Genes Dev. 2002;16(15):1964–76. https://doi.org/10.1101/gad.987402.

  19. Nagel AC, Krejci A, Tenin G, Bravo-Patiño A, Bray S, Maier D, Preiss A. Hairless-mediated repression of Notch target genes requires the combined activity of Groucho and CtBP corepressors. Mol Cell Biol. 2005;25(23):10433–41. https://doi.org/10.1128/MCB.25.23.10433-10441.2005.

  20. Oswald F, Kostezka U, Astrahantseff K, Bourteele S, Dillinger K, Zechner U, Ludwig L, Wilda M, Hameister H, Knöchel W, Liptay S, Schmid RM. SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway. EMBO J. 2002;21(20):5417–26. https://doi.org/10.1093/emboj/cdf549.

  21. Kuroda K, Han H, Tani S, Tanigaki K, Tun T, Furukawa T, Taniguchi Y, Kurooka H, Hamada Y, Toyokuni S, Honjo T. Regulation of marginal zone B cell development by MINT, a suppressor of Notch/RBP-J signaling pathway. Immunity. 2003;18(2):301–12. https://doi.org/10.1016/S1074-7613(03)00029-3.

  22. Yuan Z, VanderWielen BD, Giaimo BD, Pan L, Collins CE, Turkiewicz A, Hein K, Oswald F, Borggrefe T, Kovall RA. Structural and functional studies of the RBPJ-SHARP complex reveal a conserved corepressor binding site. Cell Rep. 2019;26:845–854.e6.

    Article  CAS  Google Scholar 

  23. Zhou S, Hayward SD. Nuclear localization of CBF1 is regulated by interactions with the SMRT corepressor complex. Mol Cell Biol. 2001;21(18):6222–32. https://doi.org/10.1128/MCB.21.18.6222-6232.2001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Fortini ME, Artavanis-Tsakonas S. The Suppressor of Hairless protein participates in Notch receptor signaling. Cell. 1994;79(2):273–82. https://doi.org/10.1016/0092-8674(94)90196-1.

  25. Furriols M, Bray S. Dissecting the mechanisms of Suppressor of Hairless function. Dev Biol. 2000;227(2):520–32. https://doi.org/10.1006/dbio.2000.9923.

  26. Maier D, Nagel AC, Johannes B, Preiss A. Subcellular localization of Hairless protein shows a major focus of activity within the nucleus. Mech Dev. 1999;89(1-2):195–9. https://doi.org/10.1016/S0925-4773(99)00208-7.

  27. Maier D, Praxenthaler H, Schulz A, Preiss A. Gain of function Notch phenotypes associated with ectopic expression of the Su(H) C-terminal domain illustrate separability of Notch and Hairless-mediated activities. PLoS One. 2013;8(11):e81578. https://doi.org/10.1371/journal.pone.0081578.

  28. Praxenthaler H, Nagel AC, Schulz A, Zimmermann M, Meier M, Schmid H, Preiss A, Maier D. Hairless-binding deficient Suppressor of Hairless alleles reveal Su(H) protein levels are dependent on complex formation with Hairless. PLoS Genet. 2017;13(5):e1006774. https://doi.org/10.1371/journal.pgen.1006774.

  29. Wolf D, Smylla TK, Reichmuth J, Hoffmeister P, Kober L, Zimmermann M, Turkiewicz A, Borggrefe T, Nagel AC, Oswald F, Preiss A, Maier D. Nucleo-cytoplasmic shuttling of Drosophila Hairless/Su(H) heterodimer as a means of regulating Notch dependent transcription. Biochim Biophys Acta Mol Cell Res. 1866;2019:1520–32.

  30. Gahr BM, Brändle F, Zimmermann M, Nagel AC. An RBPJ-Drosophila model reveals the dependence of RBPJ protein stability on the formation of transcription-regulator complexes. Cells. 2019;8(10):1252. https://doi.org/10.3390/cells8101252.

    Article  CAS  PubMed Central  Google Scholar 

  31. Brand AH, Perrimon N. Targeted gene expression as a means of altering cell fates and generating dominant phenotypes. Development. 1993;118(2):401–15.

    CAS  PubMed  Google Scholar 

  32. Roy S, Shashidhara LS, VijayRaghavan K. Muscles in the Drosophila second thoracic segment are patterned independently of autonomous homeotic gene function. Curr Biol. 1997;7(4):222–7. https://doi.org/10.1016/S0960-9822(06)00117-5.

    Article  CAS  PubMed  Google Scholar 

  33. Wacker SA, Alvarado C, von Wichert G, Knippschild U, Wiedenmann J, Clauss K, Nienhaus GU, Hameister H, Baumann B, Borggrefe T, Knöchel W, Oswald F. RITA, a novel modulator of Notch signalling, acts via nuclear export of RBP-J. EMBO J. 2011;30(1):43–56. https://doi.org/10.1038/emboj.2010.289.

  34. Steinhäuser K, Klöble P, Kreis NN, Ritter A, Friemel A, Roth S, Reichel JM, Michaelis J, Rieger MA, Louwen F, Oswald F, Yuan J. Deficiency of RITA results in multiple mitotic defects by affecting microtubule dynamics. Oncogene. 2017;36(15):2146–59. https://doi.org/10.1038/onc.2016.372.

    Article  CAS  PubMed  Google Scholar 

  35. Hoock SC, Ritter A, Steinhäuser K, Roth S, Behrends C, Oswald F, Solbach C, Louwen F, Kreis NN, Yuan J. RITA modulates cell migration and invasion by affecting focal adhesion dynamics. Mol Oncol. 2019;13(10):2121–41. https://doi.org/10.1002/1878-0261.12551.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Brockmann B, Mastel H, Oswald F, Maier D. Analysis of the interaction between human RITA and Drosophila Suppressor of Hairless. Hereditas. 2014;151(6):209–19. https://doi.org/10.1111/hrd2.00074.

  37. Maier D, Nagel AC, Preiss A. Genetic interactions between Protein Kinase D and Lobe mutants during eye development of Drosophila melanogaster. Hereditas. 2019;156(1):37. https://doi.org/10.1186/s41065-019-0113-9.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Nagel AC, Fischer P, Szawinski J, La Rosa MK, Preiss A. Cyclin G is involved in meiotic recombination repair in Drosophila melanogaster. J Cell Sci. 2012;125(22):5555–63. https://doi.org/10.1242/jcs.113902.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

We thank M. Ketelhut for expert technical assistance. We are indebted to A. Preiss for helpful discussions and input to the manuscript.

Funding

This work was funded by DFG grants (http://www.dfg.de) MA1328/11–1 to DM and NA 427/5–1 to ACN. The funding body had no role in the design of the study and collection, analysis, and interpretation of data and in writing the manuscript. Open Access funding enabled and organized by Projekt DEAL.

Author information

Authors and Affiliations

Authors

Contributions

DM and ACN conceived of the work. DW performed the histological examination of the specimen. DW, DM and ACN analyzed, interpreted and validated the data. DM and ACN supervised the work and provided resources and funding. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Anja C. Nagel.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wolf, D.B., Maier, D. & Nagel, A.C. Nucleo-cytoplasmic shuttling of murine RBPJ by Hairless protein matches that of Su(H) protein in the model system Drosophila melanogaster. Hereditas 158, 11 (2021). https://doi.org/10.1186/s41065-021-00175-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s41065-021-00175-z

Keywords