Abstract
Notch signaling is involved in the development of almost all organ systems and required post-developmentally to modulate tissue homeostasis. Rare variants in Notch signaling pathway genes are found in patients with rare Mendelian disorders, while unique or recurrent mutations in a similar set of genes are identified in cancer. The human genome contains four genes that encode Notch receptors, NOTCH1–4, all of which are linked to rare diseases and cancer. Although some mutations have been classified as clear loss- or gain-of-function alleles based on cellular or rodent based assay systems, the functional consequence of many variants/mutations in human Notch receptors remain unknown. In this review, I will first provide an overview of the domain structure of Notch receptors and discuss how each module is known to regulate Notch signaling activity in vivo using the Drosophila Notch receptor as an example. Next, I will introduce some interesting mutant alleles that have been isolated in the fly Notch gene over the past >100 years of research and discuss how studies of these mutations have facilitated the understanding of Notch biology. By identifying unique alleles of the fly Notch gene through forward genetic screens, mapping their molecular lesions and characterizing their phenotypes in depth, one can begin to unravel new mechanistic insights into how different domains of Notch fine-tune signaling output. Such information can be useful in deciphering the functional consequences of rare variants/mutations in human Notch receptors, which in turn can influence disease management and therapy.
Keywords: Notch signaling pathway, structure function analysis, Drosophila melanogaster, Mendelian diseases, cancer, variants of unknown significance (VUS)
Graphical Abstract
Missense alleles in the fly Notch gene isolated over the past >100 years of research have facilitated the understanding of Notch biology. By identifying unique alleles of the fly Notch gene through forward genetic screens, mapping their molecular lesions and characterizing their phenotypes in depth, one can begin to unravel new mechanistic insights into how different domains of Notch fine-tune signaling output. Such information can be useful in deciphering the functional consequences of rare variants/mutations in human Notch receptors, which in turn can influence disease management and therapy.
1. Unique aspects of the Notch signaling pathway
Notch signaling is conserved in most metazoan species and regulates many developmental decisions (Artavanis-Tsakonas and Muskavitch 2010; Yamamoto et al. 2014b). More recently, Notch signaling has been shown to be required in many post-developmental contexts, including tissue homeostasis (Bigas and Porcheri 2018) and neural physiology (Ables et al. 2011). Several features make the canonical Notch signaling pathway unique compared to other evolutionarily conserved core developmental signaling pathways, such as Wnt, Hedgehog, and BMP/TGF-β (Bier 2005; Salazar and Yamamoto 2018). For example, while many signaling pathways transmit signals to remotely located cells (paracrine or endocrine signaling), Notch signaling typically occurs between two juxtaposed cells (Hoppe and Greenspan 1986). One reason for the juxtacrine nature of this pathway is that both ligands [Delta or Serrate in Drosophila, Delta-like (Dll) and Jagged (Jag) family ligands in mammals] and Notch receptors are type-I transmembrane proteins that are anchored to the plasma membrane or found in other membranous compartments within the cell (Wharton et al. 1985a; Kopczynski et al. 1988; Fleming et al. 1990). In addition, unlike many cell-cell communication pathways in which the physical interaction between ligands and receptors is sufficient to trigger signal activation, Notch signal activation further requires several additional events that take place between the neighboring signal sending and receiving cells such as endocytosis that takes place in the signal sending cell (Parks et al. 2000; Nichols et al. 2007). Sequence of events following the ligand-receptor interaction ultimately causes the release of the intracellular domain of Notch (NICD) from the plasma membrane (Kopan and Ilagan 2009). Nuclear localization signals that are present in the NICD target this fragment to the nucleus where it forms a transcriptional activation complex with CSL (CBF-1, Suppressor of Hairless, Lag-2) transcription factors and Mastermind-family coactivators (Kovall and Blacklow 2010). Basic helix-loop-helix (bHLH) transcriptional repressors of the Enhancer of split [E(spl)] family proteins are representative examples of Notch signaling target genes that are evolutionarily conserved (Bray 2006). To prevent constitutive activation of the pathway, NICD becomes phosphorylated and polyubiquitinated for proteasomal degradation through its carboxyl-terminal PEST (proline, glutamic acid, serine, and threonine-rich) domain (Bray and Gomez-Lamarca 2018).
Another unique feature of Notch signaling is its stringent dosage sensitivity. In Drosophila, the Notch gene encodes the sole Notch receptor for this organism (Wharton et al. 1985a; Kopan and Ilagan 2009). Removal of one copy of Notch using a deletion or addition of an extra copy of Notch using a duplication both cause morphological alterations (Artavanis-Tsakonas and Muskavitch 2010). Similarly in mice that have four Notch paralogs (Notch1–4), Notch1 (Zhang et al. 2000; Koenig et al. 2017) and Notch2 (Witt et al. 2003) are also known to be haploinsufficient loci. Furthermore, haploinsufficiency of NOTCH1 in humans has been linked to Adams-Oliver syndrome (OMIM #616028) (Stittrich et al. 2014) and aortic valve diseases (OMIM #109730) (Garg et al. 2005). Considering that the amount of NICD that actively engages in transcription is tightly coupled to the amount of Notch receptors that are synthesized, activated and degraded due to the lack of any signal amplification steps (Salazar and Yamamoto 2018), even a modest reduction in the amount of receptors can have a significant impact on signal output.
2. Variants and mutations in human NOTCH genes that are linked to rare Mendelian diseases and cancer
Of the four genes encoding Notch receptors in the human genome, variants in NOTCH1, NOTCH2 and NOTCH3 have been associated with rare Mendelian diseases (Louvi and Artavanis-Tsakonas 2012; Salazar and Yamamoto 2018), whereas single nucleotide polymorphisms in NOTCH4 have been associated with increased risk of schizophrenia (Wei and Hemmings 2000) and systemic sclerosis (Cardinale et al. 2016; Zhou et al. 2019) in certain populations (Table 1). Some disease associated human NOTCH variants have been experimentally categorized as pathogenic loss- (LOF) or gain-of-function (GOF) alleles. For example, de novo or dominantly inherited missense or early frameshift variants in NOTCH2 that are linked to Alagille syndrome 2 (OMIM #610205) behave as LOF alleles (McDaniell et al. 2006). In contrast, late nonsense and frameshift alterations in the same gene are linked to Hajdu-Cheney syndrome (OMIM #102500), and these variants behave as GOF alleles because they result in formation of functional and more stable NICDs that lack the PEST domain (Isidor et al. 2011; Simpson et al. 2011). Interestingly in some cases, the functional consequences of disease-linked missense variants in NOTCH genes are still under debate. For example, a few variants in NOTCH3 linked to CADASIL (cerebral autosomal dominant arteriopathy with subcortical infarct and leukoencephalopathy, OMIM #125310) have been reported to behave as hypomorphic (partial LOF) alleles, while others have been reported to have no effect on Notch signaling (Joutel et al. 1996; Rutten et al. 2014). As whole-exome (WES) and whole-genome sequencing (WGS) become part of the general medical practice in the era of personalized medicine (Posey et al. 2019), the number of variants of unknown significance (VUS) in NOTCH and related genes will likely increase at a rapid pace. Understanding the precise mechanism of how each mutation affects protein function is critical since different therapeutic strategies may need to be employed based on the specific type of variant an individual carries (Wangler et al. 2017; Bellen et al. 2019; Harnish et al. 2019).
Table 1.
Gene | Disease Study | Disease Name | OMIM# | Types of variants/mutations | Proposed mechanism | Representative References |
---|---|---|---|---|---|---|
NOTCH1 | Mendelian diseases | Adams-Oliver syndrome 5 | #616028 | early truncation or missense | LOF (haploinsufficiency) | (Stittrich et al., 2014) |
Aortic valve disease 1 | #109730 | early truncation or missense | LOF (haploinsufficiency) | (Garg et al., 2005) | ||
Cancer | T-cell acute lymphoblastic leukemia (T-ALL) | - | missense in the NRR and late truncation | GOF (oncogene) | (Weng et al., 2004) | |
B-cell Chronic Lymphocytic Leukemia (B-CLL) | - | late truncation | GOF (oncogene) | (Sportoletti et al., 2010) | ||
Squamous cell carcinoma | - | early truncation, in frame deletions or missense | LOF (tumor suppressor) | (Agrawal et al., 2011; Wang et al., 2011) | ||
Bladder transitional cell carcinoma | - | early truncation, in frame deletions or missense | LOF (tumor suppressor) | (Rampias et al., 2014) | ||
Small-cell lung carcinoma | - | early truncation or missense | LOF (tumor suppressor) | (George et al., 2015) | ||
Lower-grade glioma | - | early truncation, in frame deletions or missense | LOF (tumor suppressor) | (Cancer Genome Atlas Research et al., 2015) | ||
NOTCH2 | Mendelian diseases | Alagille syndrome 2 | #610205 | early truncation, missense | LOF (haploinsufficiency) | (McDaniell et al., 2006) |
Hajdu-Cheney syndrome | #102500 | late truncation | GOF (hypermorphic) | (Isidor et al., 2011; Simpson et al., 2011) | ||
Cancer | Splenic marginal zone lymphoma | - | late truncation, missense | GOF (oncogene) | (Kiel et al. 2012; Rossi et al. 2012) | |
Chronic myelomonocytic leukaemia | - | missense | LOF (tumor suppressor) | (Klinakis et al., 2011) | ||
Bladder transitional cell carcinoma | - | early truncation, in frame deletions or missense variants | LOF (tumor suppressor) | (Rampias et al., 2014) | ||
Small-cell lung carcinoma | - | early truncation or missense variants | LOF (tumor suppressor) | (George et al., 2015) | ||
Squamous cell carcinoma | - | early truncation, missense variants | LOF (tumor suppressor) | (Wang et al., 2011) | ||
NOTCH3 | Mendelian diseases | Cerebral arteriopathy with subcortical infarcts and leukoencephalopathy 1 (CADASIL) | #125310 | missense in EGFr affecting cysteine residues (majority) | ? (neomorphic? hypomorphic?) | (Joutel et al., 1996) |
Lateral meningocele syndrome | #130720 | missense | GOF? (hypermorphic?) | (Gripp et al., 2015) | ||
Infantile myofibromatosis 2 | #615293 | late truncation | GOF (hypermorphic) | (Martignetti et al., 2013) | ||
Cancer | Bladder transitional cell carcinoma | - | early truncation or missense variants | LOF (tumor suppressor) | (Rampias et al., 2014) | |
Squamous cell carcinoma | - | early truncation, in frame deletions or missense variants | LOF (tumor suppressor) | (Wang et al., 2011) | ||
NOTCH4 | GWAS | Susceptibility to schizophrenia | - | single nucleotide polymorphisms | ? | (Wei and Hemmings, 2000) |
Susceptibility to systemic sclerosis | - | single nucleotide polymorphisms (missense) | ? | (Cardinale et al., 2016) | ||
Cancer | Squamous cell carcinoma | - | early truncation, missense | GOF? (oncogene?) | (Wang et al., 2011) |
In addition to the involvement in a handful of genetic diseases, somatic mutations in human NOTCH genes as well as alterations in their expression are found in different types of cancer (Nowell and Radtke 2017; McCarter et al. 2018; Aster et al. 2017). Similar to variants linked to rare diseases, in vitro and in vivo studies of cancer-associated NOTCH alleles have led to the classification of some mutations as LOF or GOF (Table 1). For example, late-truncation mutations in the NOTCH1 gene that remove the PEST domain are often found together with missense mutations that affect the LNR (Lin-12/Notch repeat) domain (see below) in T-cell acute lymphoblastic leukemia (T-ALL) (Weng et al. 2004). These two types of mutations act synergistically to activate Notch signaling in a ligand-independent manner to produce a NICD that is highly stable (Aster et al. 2009). Hence, these are GOF alleles and NOTCH1 acts as an oncogene in this context. Similar late-truncating and missense GOF mutations in NOTCH1 and NOTCH2 are also frequently found in B-cell chronic lymphocytic leukemia (B-CLL) (Sportoletti et al. 2010; Fabbri et al. 2011; Puente et al. 2011) and splenic marginal zone lymphoma (Kiel et al. 2012; Rossi et al. 2012), respectively. On the contrary, LOF mutations in NOTCH genes are frequently found in chronic myelomonocytic leukemia (Klinakis et al. 2011), bladder transitional cell carcinoma (Rampias et al. 2014), small-cell lung carcinoma (George et al. 2015), lower-grade glioma (Cancer Genome Atlas Research et al. 2015), and squamous cell carcinoma of the head, neck, skin and lung (Agrawal et al. 2011; Wang et al. 2011), indicating that Notch signaling acts as a tumor suppressor in these tissues. Therefore, while pharmacological strategies to inhibit Notch activity may be useful to treat certain types of leukemia, drugs that activate Notch signaling may be effective against other types of leukemia and specific types of solid tumors. Notch signaling is highly context specific, and can even perform completely opposite tasks in different tissues, cell types, and developmental stages (e.g. promoting growth in one context while inducing cell death in another). Thus in order to facilitate precision medicine, it is critical to understand the functional consequences of diverse NOTCH receptor mutations identified from cancer tissue sequencing efforts to design appropriate treatment strategies (Letai 2017).
3. Domain structure of Notch receptors and their known function
Notch receptors are composed of relatively large extracellular and intracellular domains and a single transmembrane domain that connects the two. The extracellular domain of the fly Notch receptor is ~180kDa in size, whereas the intracellular domain is ~120kDa. A number of evolutionarily conserved domains and features are found along the entire length of this ~300kDa protein (Figure 1). Here, I highlight each functional unit within the fly Notch receptor starting from the amino-terminal of the protein and discuss their roles in signal transduction.
i). Epidermal growth factor-like repeats (EGFr)
The amino acid sequence of the Drosophila Notch protein starts with a signal sequence [52 amino acids (a.a.) according to UniProt (The UniProt 2017), https://www.uniprot.org/uniprot/Q32UW5] followed by 36 epidermal growth factor-like repeats (EGFr). Human NOTCH1 and NOTCH2 also contain 36 EGFr but NOTCH3 and NOTCH4 have fewer EGFr: 34 in NOTCH3 and 29 in NOTCH4, respectively (Kopan and Ilagan 2009). Each EGFr is approximately 40 a.a. and contains six cysteine residues that form three disulfide bonds to stabilize their modular structures (Stenflo 1991). Some EGFr contain Ca+2 binding motifs at their amino-terminus and these are specifically referred to as calcium binding EGFr (cbEGFr) (Handford et al. 1991). When Ca+2 is bound to cbEGFr, it allows the EGFr to be packed into a rigid, rod-like conformation, reducing the flexibility between the cbEGFr and the preceding EGFr (Hambleton et al. 2004). In Drosophila Notch, 21 EGFr are predicted to be cbEGFr and the positions of these cbEGFr are highly conserved across evolution (Jafar-Nejad et al. 2010; Rana and Haltiwanger 2011), suggesting this pattern may help Notch receptors adopt a certain three-dimensional conformation in vivo.
Many EGFr are also subjected to post-translational O-glycosylation on the hydroxyl “O” of certain serine (S) or threonine (T) residues (Stanley and Okajima 2010). EGFr that are subjected to O-fucosylation carry a consensus sequence motif, C2-X-X-X-A/G/S-(S/T)-C3 (C2 and C3 indicate the 2nd and 3rd cysteine of an EGFr, X can be any amino acid; A, G, S, T are alanine, glycine, serine and threonine, respectively; O-fucosylation occurs on the S/T), which is the target site for the protein O-fucosyltransferase 1 (O-fut1). In Drosophila Notch, 22 EGFr are predicted to contain the O-fucosylation motif and most of these sites are indeed O-fucosylated at high stoichiometry in cultured cells (Harvey et al. 2016). O-fucosylation of Notch takes place in the endoplasmic reticulum (ER) (Okajima et al. 2005), and O-fut1 can only recognize properly folded EGFr (Wang and Spellman 1998), serving as one of the quality control mechanisms during Notch biosynthesis to ensure the production of Notch proteins with the correct structural conformation. LOF mutations or RNAi mediated knockdown of O-fut1 cause strong loss of Notch signaling phenotypes in a variety of contexts (Okajima and Irvine 2002; Sasamura et al. 2003). In O-fut1 mutant cells, Notch accumulates in the ER and fails to be exocytosed. Since this trafficking defect can be rescued by over-expression of an enzymatically inactive form of O-fut1, this protein is thought to also function as a molecular chaperone that assists in the proper folding of Notch in the ER in addition to its catalytic role as a protein glycosyltransferase (Okajima et al. 2005; Okajima et al. 2008). However, since several reports in flies and mice suggest that Notch can be delivered to the cell surface in the absence of O-fut1 (Pofut1 in mouse), the role of O-fut1 and its chaperone function is still under debate (Okamura and Saga 2008; Stahl et al. 2008; Yao et al. 2011; Sasamura et al. 2007; Sasaki et al. 2007). Furthermore, by analyzing a catalytically dead O-fut1 allele generated through a gene knock-in methodology in Drosophila, Notch receptors that lack O-fucosylation were shown to exhibit defects in endocytic trafficking at high temperature (30°C) (Ishio et al. 2015). Together with O-glucosylation that takes place on specific EGFr (see below), monosaccharide modification of Notch on multiple EGFr seems to function as a mechanism to ensure proper Notch signaling activity under certain conditions. Considering that the role of endocytic trafficking has been proposed to function as a buffering mechanism to provide robustness to Notch signaling under different temperature conditions (Shimizu et al. 2014), it would be interesting to see how changes in the structure of the extracellular domain of Notch impacts endocytic trafficking.
O-fucosylation on EGFr of Notch mediated by O-fut1 can be further elongated in the Golgi apparatus by Fringe (Fng), a β−1,3-N-acetylglucosamintransferase that catalyzes the addition of N-acetylglucosamine (GlcNAc) to O-fucosylated EGFr (Munro and Freeman 2000; Moloney et al. 2000a; Bruckner et al. 2000). The primary function of Fng is to change the affinity of the interactions between Notch and its ligands. When Notch is not Fng modified, it can bind strongly to Serrate but weakly to Delta. On the contrary, when Notch becomes Fng modified, this ligand selectivity becomes reversed and Notch can now bind strongly to Delta but only weakly to Serrate. In vitro experiments suggest that multiple EGFr that are modified by O-fut1/Fng contribute to ligand-receptor interactions and ligand selectivity (Xu et al. 2007; Harvey et al. 2016). In addition to studies based on genetics and biochemistry, recent crystallography and nuclear magnetic resonance (NMR) spectroscopy based structural biological approaches are beginning to unravel how Notch molecularly distinguishes Delta family ligands from Serrate/Jagged family ligands (Baron 2017; Handford et al. 2018), which will be discussed later in this article.
In addition to the O-fut1 mediated O-fucosylation and subsequent elongation by Fng, some EGFr of Notch can be covalently modified by O-glucose and xylose. EGFr that are O-glucosylated carry a motif, C1-X-S-X-P-C2 (C1 and C2 indicate the first and second cysteine of an EGFr, X can be any amino acid; S, P are serine and proline, respectively; O-glucosylation occurs on the S), that is recognized by Rumi, a protein O-glucosyltransferase located in the ER (Harvey and Haltiwanger 2018). Mutant alleles of rumi were originally isolated as temperature sensitive mutations in Drosophila that show severe Notch LOF phenotypes through a forward genetic screen (Acar et al. 2008). O-glucosylation of Notch is dispensable for ligand-receptor interactions but essential for S2 (site-2) cleavage (see below) of Notch at restrictive temperatures (above 25°C). At permissive temperatures (below 18°C), however, S2 cleavage and signal activation occurs normally, suggesting that alterations in thermodynamics may affect local or overall structure of the extracellular domain of Notch that is required for proper processing by ADAM proteases (Kuzbanian in Drosophila, ADAM10 in human) that mediates this cleavage. In mice, Poglut1 (rumi ortholog) knockout animals are embryonic lethal and exhibit a number of Notch signaling related developmental defects (Fernandez-Valdivia et al. 2011). In Drosophila Notch, 18 EGFr carry the consensus O-glucosylation motif and these sites are indeed efficiently modified when tested in cultured cells (Jafar-Nejad et al. 2010; Takeuchi and Haltiwanger 2010; Harvey et al. 2016). Through elegant structure-function studies performed in vivo in Drosophila using bacterial artificial chromosome (BAC) based genomic rescue constructs, O-glucosylated residues were found to contribute to robust signal activation at high temperature in an additive fashion, and no single site was found to be solely responsible for the rumi mutant phenotype (Leonardi et al. 2011).
Studies in mammalian cells as well as Drosophila have shown that the O-glucose that is added by Rumi can be further elongated by two contiguous α−1,3-linked xylose residues (Bakker et al. 2009; Moloney et al. 2000b). The addition of a xylose to glucose is mediated by a UDP-D-xylose:β-D-glucoside α−1,3-D-xylosyltransferase that is encoded by the shams gene in Drosophila (Lee et al. 2013). shams negatively regulates Notch signaling by modulating trafficking and ligand-receptor interaction of the Notch receptor (Lee et al. 2013; Lee et al. 2017). The first xylose added by Shams can be further extended by the Xyloside xylosyltransferase (Xxylt) (Pandey et al. 2018). While no obvious morphological defects are detected in Xxylt null animals, these mutants can genetically suppress the dominant haploinsufficient wing notching phenotype seen in heterozygous Notch null alleles, suggesting a modulatory role in Notch signaling that can be revealed in a sensitized genetic background (Pandey et al. 2018). Furthermore, studies on fly mutants defective in UDP-xylose (source of xylose added to EGFr) synthesis indicated that xylosylation and O-fucosylation have redundant roles in regulating Notch trafficking (Matsumoto et al. 2016), indicating that O-glycosylation of Notch has numerous unique as well as overlapping roles in fine-tuning Notch signaling. To further complicate things, biochemical studies have shown that both Drosophila and mammalian Rumi can act as a protein O-xylosyltransferase to mediate the addition of O-xylose instead of O-glucose on EGFr, establishing a dual substrate specificity (Takeuchi et al. 2011). Additionally in mammals, two Rumi-related proteins, Poglut2 and Poglut3, have been identified to O-glucosylate specific EGFr in NOTCH1 and NOTCH3 independent of Rumi/Poglut1 (Takeuchi et al. 2018). Although the function of the fly homolog of mouse Poglut2/3 encoded by the CG31139 gene has yet to be characterized, these and other studies reveal the intricate and complicated nature of post-translational modification of the extracellular domain of Notch receptors.
ii). LIN-12/Notch repeat (LNR) and heterodimerization (HD) domains
The LNR domain consists of three LNR motifs (Figure 1) (Lovendahl et al. 2018). Each LNR motif consist of a ~40 amino acid module that contains six cysteine residues that form three intramolecular disulfide bonds and a Ca2+ binding motif, similar to cbEGFr (Aster et al. 1999; Vardar et al. 2003). This evolutionarily conserved LNR domain is always present in the extracellular domain immediately following the long stretch of EGFr in all Notch receptors studied to date (Kopan and Ilagan 2009). Other than a few genes in the pappalysin family of metalloproteinases that contain LNR-like motifs (Overgaard et al. 2003), LNR motifs are uniquely found in Notch receptor family proteins, suggesting that this structure may have evolved to carry out a very specific function in Notch signaling regulation. The region between the LNR domain and the transmembrane domain is also highly conserved across evolution and is referred to as the HD domain. The HD domain is cleaved [S1 (site-1) cleavage] by Furin or Furin-like proteases during the biosynthesis process of Notch receptors in the Golgi apparatus (Logeat et al. 1998; Lake et al. 2009). The two pieces of Notch cleaved at the S1 site are still held together by strong noncovalent bonds within the HD domain, hence forming a “heterodimer” (Blaumueller et al. 1997). The three tandem LNR motifs together with the subsequent HD domain are often referred to as the negative regulatory region (NRR) because this structure protects the S2 cleavage site, located within the HD domain, from ectopic cleavage by ADAM proteases in the absence of the ligand (Gordon et al. 2007). Upon ligand binding and trans-endocytosis of the Notch extracellular domain into the signal sending cell, a mechanical pulling force opens up the NRR by dissociating the LNR domain from the HD domain (Lovendahl et al. 2018). This conformational change allows ADAM proteases to access the HD domain and execute the S2 cleavage. The extracellular domain of S1/S2 cleaved Notch is co-endocytosed into the signal sending cell together with the ligand and becomes degraded through the lysosomal pathway (D’Souza et al. 2010; Yamamoto et al. 2010). The remaining membrane bound S2 cleaved Notch is often referred to as NEXT (Notch extracellular truncation) and contains a very short extracellular domain, the transmembrane domain (TMD) and NICD. NEXT is subjected to subsequent S3 (site-3) cleavage by the γ-secretase complex (Levitan and Greenwald 1995; De Strooper et al. 1999; Struhl and Greenwald 1999; Ye et al. 1999). Ca2+ in the LNR domains also contributes to the stabilization of the NRR because chelation of extracellular Ca2+ leads to ligand-independent activation of Notch signaling in cultured cells (Rand et al. 2000).
iii). Transmembrane Domain (TMD)
The TMD of Drosophila Notch is composed of a short (21 a.a. according to UniProt) stretch of hydrophobic amino acids and the S3 cleavage site is located within the TMD proximal to the intracellular domain (Figure 1) (Kopan and Ilagan 2009). NEXT is recognized by γ-secretase, a multiprotein protease complex that mediates the intra-membrane cleavage of Notch and other transmembrane proteins (Jorissen and De Strooper 2010). Following the S3 cleavage, γ-secretase performs an additional site-4 (S4) cleavage within the TMD proximal to the extracellular domain, generating small ~20 residue peptides called Nβ (Okochi et al. 2002). This is analogous to Aβ (amyloid-β) peptides generated from amyloid precursor protein (APP) that aggregate and accumulate as amyloid plaques in brains of Alzheimer’s disease patients (Selkoe and Hardy 2016). Interestingly, in vitro data suggest that familial Alzheimer disease linked variants in Presenilin genes, encoding the catalytic subunits of the γ-secretase complex, cause a shift not only in γ-cleavage site of APP but also the S4 cleavage site of Notch to generate Nβ peptides that are longer than usual (Okochi et al. 2006). However, the in vivo significance of this S4 cleavage and Nβ peptide is still unknown.
iv). Notch Intracellular Domain (NICD)
NICD consists of a single RAM (RBP-jκ Associated Molecule) domain, seven ankyrin (ANK) repeats, a transactivation domain (TAD) and a PEST sequence (Figure 1) (Bray and Gomez-Lamarca 2018). In Drosophila, a poly-glutamine (Q)-rich domain called the OPA domain is present within the TAD (Wharton et al. 1985b) but this feature is not shared among mammalian Notch receptors. In addition, NICD carries multiple nuclear localization sequences (NLS) (Kopan et al. 1994), as well as target sites for multiple post-translational modifications including phosphorylation (Espinosa et al. 2003; Ranganathan et al. 2011), acetylation (Guarani et al. 2011; Palermo et al. 2012), methylation (Hein et al. 2015), and ubiquitination (Fryer et al. 2004). NICD released from the plasma membrane translocates into the nucleus via the NLS mediated nuclear import machinery (Huenniger et al. 2010; Sachan et al. 2013).
In the nucleus, fly NICD interacts with Suppressor of Hairless [Su(H)], a DNA binding transcription factor, through the RAM and ANK domains (Kovall and Hendrickson 2004; Bray 2006). Su(H) and its orthologs are often referred to as CSL (CBF1/Su(H)/LAG-1) proteins and the human CSL is encoded by the RBPJ (Recombination signal binding protein for immunoglobulin kappa J region) gene (Tanigaki and Honjo 2010). CSL proteins bind to consensus DNA sequences (5’-TGGGAA-3’) in both the absence or presence of NICD. In the absence of NICD, Su(H) that is bound to DNA recruits transcriptional co-repressors such as Hairless, CtBP (C-terminal Binding Protein) and Groucho, which in turn recruit histone deacetylases (HDACs) and other repressive chromatin regulators to negatively regulate the expression of Notch target genes (Nagel et al. 2005; Morel et al. 2001; Kao et al. 1998; Oswald et al. 2005). Upon Notch signal activation and NICD binding to Su(H), the co-repressor complex disassembles, leading to de-repression of target gene transcription (Morel and Schweisguth 2000; Barolo et al. 2002). Mastermind, a co-activator that interacts with both NICD and Su(H), is also recruited to the complex to form a tripartite complex (Nam et al. 2006; Wilson and Kovall 2006). NICD, Su(H) and Mastermind together recruit histone acetyltransferases (HATs) such as p300 and other chromatin remodeling complexes to actively promote gene transcription and expression of Notch target genes (Kurooka and Honjo 2000; Oswald et al. 2001; Fryer et al. 2002). Signal termination is mediated by phosphorylation and subsequent ubiquitin proteasome pathway mediated degradation of NICD. Although this process has not been extensively studied in Drosophila, mammalian studies suggest that phosphorylation of NICD by kinases such as CDK8 (Cyclin-dependent kinase 8) (Fryer et al. 2004) and poly-ubiquitination by E3 ubiquitin ligases such as FBXW7 (F-Box And WD Repeat Domain Containing 7) (Oberg et al. 2001; Wu et al. 2001; Gupta-Rossi et al. 2001) facilitate the degradation of NICD in the proteasome (Schweisguth 1999; Kopan 1999).
4. Unique missense mutations in the fly Notch gene
Since the isolation of the first female fly with a notched wing phenotype in 1913 by John S. Dexter in Thomas Hunt Morgan’s laboratory (Dexter 1914), over 350 alleles of Notch (official gene symbol: N) have been reported to date in FlyBase (http://flybase.org/reports/FBgn0004647#alleles_main). Other than a few mutants such as the original N allele that appeared spontaneously while housing a large number of flies in culture (Morgan and Bridges 1916), most alleles were obtained through diverse mutagenesis techniques including physical mutagenesis using X-rays and γ-rays, chemical mutagenesis using EMS (ethyl methanesulfonate) and ENU (ethyl nitrosourea), and mobilization of transposable elements such as P-elements and piggybac (Lindsley and Zimm 1992). Surprisingly, the molecular identity of majority of reported Notch alleles are still unknown because most were identified and characterized based on their phenotypes and basic genetic experiments such as complementation tests and rescue experiments. Many mutants including N55e11 and N54l9 [also known as Df(1)N-54l9] that have been frequently used in genetic experiments behave as null alleles (amorph) (Kidd et al. 1983; Mohler 1956). Amorphic N alleles were critical for the mapping and cloning of the Notch gene in 1985 (Wharton et al. 1985a), a key event in Notch signaling research history that catalyzed the dramatic expansion of the field in the following three decades. Amorphic Notch alleles in Drosophila are recessive lethal during embryogenesis and exhibit numerous developmental defects including a characteristic “neurogenic” phenotype that leads to epidermal-to-neuronal transformation during cell fate specification of the ectoderm (Poulson 1937; Lehmann et al. 1981). In fact, this neurogenic phenotype of N became the first bridge between two academic fields that were considered to be incompatible with one another at the time: genetics and developmental biology (Artavanis-Tsakonas and Muskavitch 2010). FLP/FRT based clonal analysis of amorphic N alleles have also revealed the involvement of Notch signaling in various post-embryonic developmental contexts (Salazar and Yamamoto 2018).
Currently (as of July 2019), 106 out of 367 Notch alleles curated in FlyBase (http://flybase.org) have been reported to have a defined molecular or cytological lesion. Out of these 106 defined Notch alleles, only 18 are missense alleles (Table 2) (Hartley et al. 1987; Kelley et al. 1987; Brennan et al. 1997). Despite this small number, genetic, phenotypic, biochemical and cell biological studies on a subset of these mutations have provided novel insights into how Notch signaling is regulated in vivo. In this section, I will discuss how studies of these missense alleles in Drosophila have improved our understanding of ligand-receptor interactions, higher-order protein complex formation and non-canonical signaling.
Table 2.
Drosophila Notch allele | FlyBase ID | Affected Domain(s) | Amino acid change(s) | Notes |
---|---|---|---|---|
nd-3 | FBal0012892 | EGFr-3 | p.C105F | Weak hypomorphic allele (affects wing) |
jigsaw | FBal0280339 | EGFr-8 | p.V361M | Separation of function allele (inhibits Notch-Serrate binding) |
M1 | FBal0029949 | EGFr-12 | p.E491V | Antimorphic allele (likely have an impact on both Notch-Delta and Notch-Serrate binding) |
spl-1 | FBal0012900 | EGFr-14 | p.I578T | Weak hypomorphic allele (affects eye and bristle), but may also be hypermorphic or antimorphic |
Mcd5 | FBal0130887 | EGFr-18 | p.C739Y | Separation of function allele (affects non-canonical signaling) |
Ax-59b | FBal0012853 | EGFr-24 | p.C972G | Complecated alleles (Abruptex alleles that have mixed features of gain- and loss-of-function) |
Ax-59d | FBal0012854 | p.C972YS | ||
Ax-9 | FBal0012851 | p.D948V | ||
Ax-1 | FBal0012850 | EGFr-25 | p.N986I | |
Ax-M1 | FBal0029940 | p.C999Y | ||
Ax-71d | FBal0028450 | EGFr-27 | p.S1088I | |
Ax-E2 | FBal0012858 | EGFr-29 | p.H1167Y | |
Ax-16 | FBal0012852 | p.G1174A | ||
ts1 | FBal0012887 | EGFr-32 | p.G1272D | Temperature sensitive hypomorphic alleles |
l1N-B | FBal0012886 | LNR3 | p.G1573V | Hypermorphic alleles |
414 | FBal0066009 | LNR3 & HD | p.D1577G & p.Q1641R | |
su42c | FBal0033873 | ANK-5 | p.A2060V | Strong hypomorphic alleles |
1081 | FBal0058512 | TAD | p.G2490C |
i. NM1: an antimorphic allele in the ligand-receptor interaction domain
For Drosophila Notch and human NOTCH proteins that have 36 EGFr (NOTCH1 and NOTCH2) in their extracellular domains, EGFr-11 and -12 are absolutely essential for ligand-receptor binding and signaling (Tien et al. 2009; Kovall and Blacklow 2010). Through cell culture based studies in Drosophila, Notch without EGFr-11/12 was shown to be completely defective in ligand-receptor binding (Rebay et al. 1991). In mice, an internal deletion of multiple EGFr including EGFr-11/12 of Notch1 (EGFr-8~12) also exhibited strong LOF phenotypes that resulted in embryonic lethality indistinguishable from Notch1 null mutant embryos (Ge et al. 2008). Genetic, biochemical, and cell biological experiments all pointed to the importance of these specific EGFr in ligand-receptor interactions, and molecular structures of some portions of Notch and its ligands were determined through crystallography (Cordle et al. 2008) and NMR (Hambleton et al. 2004). However, the actual structural configuration of Notch receptors bound to its ligands was not solved until very recently due to the large size of the proteins, transmembrane nature of both ligands and receptors, and relatively weak molecular interactions that had been difficult to capture. In 2015, the first co-crystal structure of a Notch receptor binding to a Delta family ligand was reported, providing direct structural evidence that EGFr-11 and EGFr-12 are indeed directly engaging in ligand-receptor interactions (Luca et al. 2015). Subsequent study by the same group further showed that EGFr-11/12 are also directly involved in the interaction between Notch receptors to Serrate/Jagged family ligands (Luca et al. 2017). Although these studies are still based on relatively small fragments of Notch and its ligands compared to the full-length proteins (EGFr-11~13 for Notch1-Dll4 complex, EGFr-8~12 for Notch1-Jag1 complex), they are indeed critical milestones towards a full structural understanding of the Notch receptor in action in vivo.
The NM1 allele was first reported in 1993 through an EMS-mediated chemical mutagenesis screen and carries a p.E491V mutation in the Ca2+ binding domain of EGFr-12 (de Celis et al. 1993; Baron 2017). Hemizygous and homozygous NM1 mutants exhibit recessive embryonic lethality and neurogenic phenotypes similar to amorphic alleles (Couso and Martinez Arias 1994; de Celis et al. 1993). Furthermore, heterozygous NM1 mutant flies also exhibit dominant morphological phenotypes in the wing and notum. Interestingly, these dominant phenotypes are stronger than amorphic alleles and cannot be fully rescued by introducing a wild-type copy of Notch, suggesting that this allele encodes a dominant negative (antimorphic) protein (de Celis et al. 1993). Although the structural consequences of this allele have not been further investigated, the fact that this mutant carries a missense change in EGFr-12 suggests that the NM1 mutant may have altered ligand binding properties. Since one would expect that a simple loss of ligand binding ability should behave as an amorphic allele similar to the Notch1 knock-in mouse that lacks the ligand binding domain (Ge et al. 2008), understanding how this allele behaves as an antimorphic allele will likely provide additional structural insights into ligand-receptor binding and signal activation.
ii. Njigsaw: a separation of function allele for ligand selectivity
Although EGFr-11 and EGFr-12 are necessary for ligand-receptor interactions, they are not fully sufficient. For example, although a small fragment of Notch containing only EGFr-11/12 was shown to be able to interact with its ligands based on cellular assays, this interaction is significantly weaker compared to the full length Notch (Xu et al. 2005). Additional experiments that deleted different domains of Notch and biochemically assessed their binding ability showed that EGFr-6~9 and EGFr-25~36 strengthen the ligand-receptor binding mediated by EGFr-11/12, while EGFr-1~5 are dispensable for or may even play a negative role in these interactions (Xu et al. 2005). In addition to some EGFr being involved in regulating both Notch-Delta and Notch-Serrate interactions, other domains within the extracellular domain were predicted to function in distinguishing the two ligand types based on data obtained from fng mutants (Irvine and Wieschaus 1994; Fleming et al. 1997; Panin et al. 1997). However, considering that Fng modifies many EGFr scattered across the protein (Harvey et al. 2016), whether a single or multiple EGFr were responsible for this function was unclear.
We reported the isolation of Njigsaw in 2012, a peculiar missense allele (p.V361M) in EGFr-8 that was obtained through a somatic mosaic chemical mutagenesis screen using EMS (Yamamoto et al. 2012). This screen was originally designed to isolate recessive lethal mutations on the fly X-chromosome that exhibit developmental or neurological phenotypes in homozygous mutant clones induced in a tissue specific manner using the FLP/FRT system (Yamamoto et al. 2014a; Deal and Yamamoto 2018). In addition to discovering a new gene that regulates Notch signaling (Charng et al. 2014b), we isolated 42 new recessive lethal alleles of Notch. Through molecular and cell biological characterization of these mutants, we found 20 of the 42 alleles were missense alleles that affect different domains of Notch (Yamamoto et al. 2012). Njigsaw was distinct from other alleles for several reasons. First, flies with homozygous Njigsaw clones exhibited strong wing phenotypes while having no effect on the development of mechanosensory bristles. Considering that flies with amorphic N clones exhibit strong phenotypes in both wings and bristles, this suggested that Njigsaw is a separation of function allele. Second, in the flies with homozygous Njigsaw clones that exhibited wing defects, we observed both loss (notching) and gain (ectopic) of wing margin tissue even within the same wing, analogous to a piece of a jigsaw puzzle. Because the former phenotype is associated with loss of Notch signaling while the latter is linked to gain of Notch signaling, this result suggested that this mutant may not be a simple LOF or GOF allele. Third, considering that a wild-type copy of Notch was able to rescue all phenotypes we observed in this mutant, we hypothesized that Njigsaw was some sort of a LOF allele that was missing a specific function.
Through series of experiments in vivo and in vitro, we demonstrated that this specific missense mutation in EGFr-8 strongly diminishes the ability of Notch to bind to Serrate without affecting its interactions with Delta. We further determined that Njigsaw was exhibiting both loss and gain of Notch signaling phenotypes in mosaic tissues because this mutation prevented both trans and cis interactions between Notch and Serrate. In addition to binding to ligands presented from neighboring cells in trans, Notch receptors can interact with ligands that are expressed in the same cell in cis (del Alamo et al. 2011; Sprinzak et al. 2010). Ligands presented in cis and trans both compete for the same ligand binding domain of the receptor (Cordle et al. 2008). Since cis-interactions cannot trigger the S2 cleavage event due to lack of mechanical tension that is provided by trans-endocytosis of the ligand, Notch signaling diminishes cell autonomously when ligands and receptors are bound in cis. This “cis-inhibition” was first documented through ligand over-expression experiments during wing development in Drosophila (Micchelli et al. 1997), and was subsequently shown to fine-tune the level of Notch signaling in vivo during the development of the fly eye (Miller et al. 2009) and oogenesis (Palmer et al. 2014).
A variation of “cis-inhibition” between Notch and Serrate (but not Delta) has been reported and referred to as “ligand cis-inhibition” (Becam et al. 2010). In this scenario, Notch receptors regulate the amount of Serrate present at the cell surface through cis-interactions and subsequent endolysosomal degradation. Since Njigsaw cannot bind to Serrate, the level of Serrate in Njigsaw mutant cells increases, which in turn mediates ectopic activation of Notch signaling in neighboring wild-type cells in a mosaic animal (Yamamoto et al. 2012). By solving this “Njigsaw puzzle”, we were able to identify the first residue outside of EGFr-11/12 in Notch that was responsible for ligand binding and discrimination, a finding that was explained structurally when the first co-crystal structure of a Notch receptor binding to a Serrate/Jagged family ligand was published five years later (Luca et al. 2017). Considering that a variant analogous to Njigsaw in mouse Notch2 (p.V327M) also exhibited defects in Jag1 mediated signaling without affecting Dll1 mediated signaling in cultured cells (Yamamoto et al. 2012), similar variants and alleles in human NOTCH orthologs may impact ligand binding and selectivity in both cis and trans.
iii). Abruptex (Ax) mutations: complicated allelic series that suggest Notch receptors functions as dimers or multimers in vivo
The region between EGFr-24 and EGFr-29 of Drosophila Notch is referred to as the Abruptex (Ax) domain (Kelley et al. 1987). This name originates from a series of missense alleles that cluster in this region, first documented in 1930 (Nasarenko 1930). All NAx mutants exhibit shared dominant phenotypes that are characterized by shortening (abrupt) of the fifth longitudinal wing vein and mild reduction of mechanosensory bristles on the dorsal thorax (Lindsley and Zimm 1992). Because these dominant phenotypes can be explained by mild increases in Notch signaling (de Celis and Garcia-Bellido 1994; Heitzler and Simpson 1993; de Celis and Bray 2000), this domain likely plays an inhibitory role in Notch signaling in certain contexts. Molecularly, the Ax domain can physically interact with a region in the extracellular domain that contains EGFr-11/12 (Pei and Baker 2008), suggesting they may mediate intra- or inter-molecular interactions within or between Notch proteins. Hence, ligands may need to physically compete with the Ax domain and overcome this inhibitory interaction to be able to bind and activate the receptor. In Notch receptors that carry Ax alleles, the interaction between the ligand binding domain and Ax domain may be lost or reduced, allowing Notch to be more easily activated. However, structural evidence is currently lacking to support this model since a crystal or cryo-electron microscopy (EM) structure of a Notch receptor that includes the Ax domain is still lacking.
It is important to note, however, that many aspects of Ax mutations cannot be explained by a simple GOF mechanism, and we still do not have a full understanding of what this domain actually does in vivo. Some Ax mutants are recessive lethal (lethal Ax), while others are homozygous viable. Homozygous viable Ax mutants can be further subdivided into two classes based on how they behave when crossed to amorphic Notch alleles. The first class of Ax alleles is called “enhancer of Notch” because transheterozygous flies (NAx/Nnull) have stronger wing notching phenotypes compared to flies heterozygous for an amorphic allele (N+/Nnull). The second class of mutants is called “suppressor of Notch” because NAx/Nnull animals exhibit milder wing notching phenotypes compared to N+/Nnull. Ax alleles that are “suppressor of Notch” are found in EGFr-24 and EGFr-25 whereas “enhancer of Notch” are located in EGFr-27 and EGFr-29 (Kelley et al. 1987), suggesting that the Ax domain can be further functionally subdivided. Surprisingly, transheterozygous (compound heterozygous) flies with one “enhancer of Notch” Ax allele and another “suppressor of Notch” Ax allele are lethal or semi-lethal (Foster 1975; Portin 1975). This extremely rare genetic phenomenon is called “negative complementation” because it is completely opposite of the well-known “complementation” phenomenon in which two recessive lethal mutants become viable when put in trans (often used as a genetic evidence that two mutations affect different genes or different functional domains within a single protein).
One molecular explanation for this mysterious “negative complementation” seen among certain allelic combinations of NAx mutants is that the true functional unit of a Notch receptor in vivo is not a monomer, but a dimer or multimer (de Celis and Garcia-Bellido 1994). Notch receptors within one class of Ax mutation may still be able to form functional complexes with themselves, but these mutants may lack compatibility with Notch receptors that carry the other class of Ax alleles. Hence, when two classes of Ax Notch receptors become expressed in a single cell in transheterozygous animals, most multi-Notch receptor containing complexes may become nonfunctional. This idea is supported by several biochemical experiments including co-immunoprecipitation (Vooijs et al. 2004), denaturing gel electrophoresis (Kidd et al. 1989) and cross-linking experiments (Sakamoto et al. 2005). Recently, a single particle EM based study visualized the extracellular domain of Notch as a dimer that can be found in different conformational states (Kelly et al. 2010). It is interesting to note that the NICD has also been shown to form a dimer when bound to paired CSL-binding sites on DNA to boost transcription of certain target genes (Arnett et al. 2010). Hence, it is important to understand the structural basis by which different domains within both extracellular and intracellular regions of Notch contribute to higher-order complex formation to obtain a full understanding of how this receptor works at the molecular level. Further structural biological studies of Notch and its ligands in their native states at atomic resolution using techniques such as cryo-EM (Yang et al. 2019) may allow scientists to observe how these proteins actually interact in vivo.
iv). Microchaetae defective (Mcd) mutations: series of alleles that implicate Notch in a non-canonical signaling pathway
A series of Microchaetae defective (Mcd) alleles of Notch exhibit a dominant phenotype that is characterized by reduction of the number of small mechanosensory bristles (microchaetae) present on the dorsal thorax of adult flies (Brennan and Gardner 2002; Heitzler 2010). NMcd alleles are also recessive lethal and form an allelic series in terms of their phenotypic strength (Ramain et al. 2001). Interestingly, all but one allele carries late truncation mutations in the intracellular domain distal to the ANK domain, deleting part of the TAD and PEST domains. Similar to late truncation variants in NOTCH2 and NOTCH3 that are linked to Hajdu-Cheney syndrome (OMIM #102500) (Isidor et al. 2011; Simpson et al. 2011) and lateral meningocele syndrome (OMIM #615293) (Gripp et al. 2015), respectively, these mutations are predicted to increase the stability of the NICD, leading to a GOF phenotype. As mentioned earlier, analogous truncation mutations in NOTCH1 are also frequently found in T-ALL (Weng et al. 2004) and B-CLL (Puente et al. 2011). Interestingly, one NMcd allele, named Mcd5 carries a p.C739Y missense mutation in one of the six cysteine residues that form the backbone of EGFr-18 (Ramain et al. 2001). Dominant phenotypes seen in NMcd5 and other NMcd truncation alleles were reported to be similar, suggesting that this missense allele somehow mimics the effect of late truncation alleles. How this may occur structurally and mechanistically has not been further investigated.
Sensory organ precursor cells (SOPs) that form mechanosensory bristles are specified from proneuronal clusters through lateral inhibition mediated by Notch, Delta, Su(H) and E(spl) (Charng et al. 2014a; Shaya and Sprinzak 2011). GOF of Notch signaling can cause a decrease in the number of SOPs by specifying more epithelial cells at the expense of SOPs during this process, which is why some bristles are missing in NAx mutants (Brennan et al. 1999). Surprisingly, however, the loss of bristle phenotype seen in NMcd alleles have been reported to be independent of Delta, Su(H), and E(spl), suggesting that NMcd alleles affect a non-canonical aspect of Notch function (Ramain et al. 2001; Alfred and Vaccari 2018). Phenotypes seen in NMcd alleles are dependent on deltex, an endocytic trafficking gene that regulates both ligand-dependent and -independent Notch signaling events (Xu and Artavanis-Tsakonas 1990; Fuwa et al. 2006). In addition, NMcd phenotypes depend on several proteins that are considered to function as core components of the canonical Wnt signaling pathway in Drosophila including dishevelled (dsh, a cytoplasmic phoshoprotein), shaggy (sgg, a cytoplasmic kinase GSK3β) and armadillo (arm, also known as β-Catenin) (Ramain et al. 2001). Although the precise mechanisms of how NMcd alleles cause the loss of bristle phenotype is unknown, yeast-two-hybrid experiments have revealed that Dsh can bind to the carboxy-terminal end of Notch, which is missing in all of the truncated NMcd alleles (Axelrod et al. 1996; Ramain et al. 2001). Therefore, NMcd alleles may be pointing us to a molecular interface where Notch signaling intersects with Wnt signaling. Cross-talk between Notch and Wnt signaling pathways have been reported in diverse developmental and pathophysiological contexts (Borggrefe et al. 2016; Collu et al. 2014). Indeed, Wingless (Wg), a homolog of mammalian WNT1 and the major ligand of the Wg/Wnt pathway in flies, has also been reported to interact with the extracellular domain of Notch to modulate its function (Wesley 1999; Couso and Martinez Arias 1994). By investigating how the domains mutated in NMcd alleles affect canonical as well as non-canonical aspects of these two signaling pathways in vivo, we will likely have a better understanding of Notch-related diseases that are caused by late truncating NOTCH alleles, such as Hajdu-Cheney syndrome, lateral meningocele syndrome, and certain forms of leukemia.
v). Nnd-3, Nspl-1 and Nts1: Additional mutations in EGFr with unique characteristics
In addition to NM1, Njigsaw, NAx and NMcd5 that I have discussed so far, several additional alleles of Notch have been reported to carry missense mutations in EGFr. These alleles also exhibit unique features, providing us with a glimpse of how pleiotropic the extracellular domain of Notch actually is. A rare genetic phenomenon referred to as “intragenic complementation”, in which two different alleles of Notch complement each other’s phenotype, can be observed between certain allelic combinations, suggesting that these mutations impact different functions within the Notch receptor. Even though many alleles were isolated in the early days of genetics, very little is known about how these mutations alter the function and structure of Notch. Phenotypic description of many classical Notch alleles can be found in “The Red Book” (Lindsley and Zimm 1992) as well as in FlyBase (Thurmond et al. 2019).
notchoid (nd) alleles of Notch are a group of recessive visible mutations that exhibit wing notching. Of these, Nnd-3 has been reported to carry a missense mutation in one of the six cysteine residues that form the core disulfide bonds in EGFr-3 (p.C105F) (Lyman and Young 1993). Since Nnd-3 is hemizygous/homozygous viable and these animals only exhibit a mild defect in wing margin (wing notching) and wing vein (thickening of the ends of wing veins, referred to as “wing delta” which is also seen upon Delta haploinsufficiency) development, it is considered as a weak hypomorphic (mild LOF) allele. If this mutation only disrupts the local structure of this specific EGFr, it is possible that EGFr-3 is dispensable for most Notch signaling processes. Interestingly, the majority of NOTCH3 variants that are linked to CADASIL are alleles that decrease (5 cysteine/EGFr) or increase (7 cysteine/EGFr) the number of cysteine residues in EGFr, similar to Nnd-3 and NMcd5 mutations (Joutel et al. 1996). Odd numbers of cysteines per EGFr leads to an unpaired cysteine that may form abnormal disulfide bonds within NOTCH3 or with other proteins that it may interact with. Although some CADASIL alleles behave as LOF alleles when tested in cells or in mice (Arboleda-Velasquez et al. 2011), other variants do not have an obvious effect on canonical Notch signaling, suggesting that this disease may be caused a neomorphic mechanism. For example, a cysteine residue in NOTCH3 that failed to form intramolecular disulfide bonds may form covalent bonds with other extracellular/luminal molecules in its vicinity, thereby modulating the function of these unidentified target proteins. In support of this idea, one key diagnostic criteria of CADASIL in addition to genetic testing is the accumulation of granular osmophilic material (GOM), an electron dense structure that contains the extracellular domain of NOTCH3 and other proteins, in vascular tissues (Baudrimont et al. 1993; Ishiko et al. 2006). A similar but recessively inherited condition called CARASIL (Cerebral autosomal recessive arteriopathy with subcortical infarcts and leukoencephalopathy, OMIM #600142) is caused by mutations in HTRA1, a gene encoding a protease involved in TGF-β signaling (Hara et al. 2009). Thus, it would be interesting to explore whether NOTCH3 with CADASIL mutations may aberrantly interact with TGF-β signaling factors or other HTRA1 substrates in order to probe the molecular mechanism of these disorders. Another interesting but unexplained aspect of CADASIL alleles is that although NOTCH3 has 34 EGFr, most of the variants linked to this disease are found in EGFr-1~5 (corresponding to EGFr-1~6 in Drosophila Notch which has 36 EGFr) (Joutel et al. 1997; Masek and Andersson 2017). Understanding the effect of cysteine mutations of these early EGFr compared to other EGFr may provide us with some hints to why such a bias exists. Hence, functional studies of Nnd-3 (EGFr-3) as well as other N alleles that affect cysteine residues in EGFr such as NMcd5 (EGFr-18) may provide useful information that benefits patients who suffer from this devastating neurodegenerative disease.
split (spl) alleles of Notch are recessive visible mutations that primarily manifest in eyes and bristles (Lindsley and Zimm 1992). Homozygous females and hemizygous males have small rough eyes and their bristles are often split, doubled or sometimes missing (Lees and Waddington 1942). Interestingly, Nspl alleles do not exhibit any wing defects on their own, and show intragenic complementation with a number of other recessive visible N mutations including Nnd and NAx alleles (Shellenbarger and Mohler 1975; Brennan et al. 1997). Nspl-1 is caused by a missense mutation in EGFr-14 (p.I578T), but the exact genetic effect of this allele is controversial. Some studies refer to this allele as a hypomorph based on complementation with strong N LOF alleles (Brennan et al. 1997), whereas others document this as a hypermorph (Hing et al. 1999) or an antimorph (de Celis and Garcia-Bellido 1994) based on other genetic tests. The reason why this single mutation can be genetically classified as different allele types based on different experiments may be due to the complicated role of this residue and domain in different developmental contexts. Although the precise function of EGFr-14 and this specific amino acid in Notch is still unknown, the study of Nspl alleles was critical for Notch research because genetic screens to identify intergenic modifiers of Nspl led to the identification of the E(spl) complex, a group of genes encoding evolutionarily conserved bHLH transcriptional repressors that function downstream of the canonical Notch signaling pathway (Von Halle 1965; Lehmann et al. 1983). Without the pioneering study of Nspl and E(spl) mutations, the mammalian HES (Hairy and Enhancer-of-Split) (Akazawa et al. 1992; Sasai et al. 1992) and HEY (Hairy/Enhancer-of-split related with YRPW motif protein) (Leimeister et al. 1999) genes that are homologous to Drosophila E(spl) genes would have been named differently.
Some Notch alleles are known to be temperature sensitive (Shellenbarger and Mohler 1975; Lindsley and Zimm 1992). This could be due to the variant causing destabilization of the protein at restrictive (high) temperatures, or it could be due to other reasons such as those revealed through studies of genes involved in O-glycosylation of Notch (see above). Nts1 (also known as Nl1N-ts1) carries a missense mutation in an O-fucosylation motif of EGFr-32 (p.G1272D) and behaves as a temperature sensitive hypomorphic allele (Xu et al. 1992). Independent mutations named Nts2 (Shellenbarger and Mohler 1975) and Nts3 (Li et al. 2009; Nicholson et al. 2011) have been reported to function in a similar manner but their molecular lesions have yet to be determined. Although the mechanisms by which these mutations inhibit Notch signaling at their restrictive temperatures are unknown, these alleles have been useful in assessing the function of Notch signaling in diverse developmental processes post-embryonically. By keeping the Nts animals at a permissive temperature (18°C) during embryogenesis, and shifting the animals to a restrictive temperature (29°C) during specific time points of interest, one can perform “conditional knockout” experiments of Notch in a semi-reversible manner. Such experiments elucidated that Notch signaling is used not only once but reiteratively during the formation of variety of tissues such as the fly eye (Cagan and Ready 1989), leg (Bishop et al. 1999), wing (Micchelli and Blair 1999), and mechanosensory bristles (Hartenstein and Posakony 1990). Furthermore, these alleles have allowed researchers to explore the post-developmental roles of Notch signaling in adult flies, revealing its roles in tissue homeostasis through regulation of stem cells (Song et al. 2007; Ohlstein and Spradling 2006; Micchelli and Perrimon 2006; Chaturvedi et al. 2017), stereotypic behavior including learning and memory (Presente et al. 2004), and longevity (Presente et al. 2001). Our understanding of Notch signal regulation will likely advance if we can precisely understand how the Nts1 missense mutation in EGFr-32 causes these temperature sensitive phenotypes at the molecular level. In addition, mapping of the Nts2 and Nts3 alleles as well as identifying other missense mutants in Notch that cause conditional phenotypes will help decipher how the Notch receptor works in different environmental settings in vivo.
5. Conclusions
In contrast to most signaling pathways that harbor the name of the major ligand in the pathway (e.g. Wnt, Hedgehog, BMP/TGF-β, EGF, FGF, IGF/Insulin, PDGF, VEGF…), Notch seems to be one of the few major developmental signaling pathways that bears the name of the receptor, implicating the central role of this protein in the entire signaling network (Guruharsha et al. 2012). Since the isolation of the first Notch allele in the early 20th century, hundreds of Notch alleles with different molecular lesions and unique phenotypic and molecular characteristics have been reported. While amorphic alleles of Notch have been critical tools to assess which biological events depend on Notch signaling, various missense alleles have provided insights into how this pleiotropic and complicated pathway is fine-tuned in vivo. In this review, I primarily focused on alleles that affect the EGFr of the Notch receptor and discussed how genetic analyses have provided insights into the workings of this molecular pathway. Several missense mutations that are found in other domains of the Notch receptor are also worth mentioning. N414 and Nl1N-B carry a p.D1577G and a p.G1573V change in the third LNR repeat, respectively (Brennan et al. 1997; Lyman and Young 1993). Similar to some of the missense mutations associated with T-ALL in NOTCH1 (Weng et al. 2004) and a variant linked to infantile myofibromatosis 2 in NOTCH3 (OMIM #615293) (Martignetti et al. 2013; Xu et al. 2015), these mutations behave as GOF alleles likely due to ligand-independent activation mediated by unfolding of the NRR. Some other interesting missense mutations are also found in the intracellular domain of Notch, including Nsu42c that carries a p.A2060V change in the fifth ANK repeat (Diederich et al. 1994) and N1081 carries a p.G2490C change within the TAD (Brennan et al. 1997). These mutations behave as strong LOF alleles, providing functional information that may be useful to understand of how NICD binds to its interaction partners in the nucleus. Finally, in addition to the NMcd truncation alleles discussed earlier in this article, there are a number of nonsense and frameshift alleles of N that exhibit unique phenotypes. By comparing of phenotypes seen in mutants such as NCo (nonsense allele within the RAM domain, deleting most of the intracellular domain) (Gottschewski 1935; Lyman and Young 1993), N60g11 (frameshift allele immediately after the ANK repeats) (Welshons et al. 1963; Brennan et al. 1997), Nl1N-3 (frameshift allele deleting PEST and part of the TAD including the OPA domain identical to NMcd3) (Welshons 1965; Brennan et al. 1997), and Nnd-2 (introduces a frameshift 14 residues before the carboxy-terminal end of the protein) (Welshons 1965; Xu et al. 1990), we can begin to understand how different domains within the NICD function in both canonical and non-canonical modes of Notch signaling. Since the intracellular domain of type-I transmembrane receptors are often required for proper intracellular trafficking (Yamamoto et al. 2010), it would be interesting to assess what aspect, if any, of the phenotypes seen in these intracellular domain mutations can be attributed to defects in exocytosis, endocytosis or endolysosomal trafficking (Schnute et al. 2018). Considering such knowledge cannot be obtained by studying amorphic alleles of Notch or through gene knockdown studies using RNAi, studies of unique missense, nonsense and frameshift mutation will continue to provide novel insights into this pathway.
Although advancement of genome editing technologies such as the CRISPR/Cas9 system are allowing scientists to introduce virtually any mutation into any gene in any species (Knott and Doudna 2018), forward genetic screens using chemical mutagens such as EMS will still be one of the most powerful ways to reveal specific residues or domains in Notch and other proteins that have unique functions in vivo (Yamamoto et al. 2012). Although the N gene has been extensively studied by numerous laboratories in the past, interesting mutations in this gene have not yet reached saturation. In addition, assessments of disease associated missense variants/mutations in human NOTCH genes may provide novel structural and functional information on how Notch receptors are regulated. If human NOTCH receptors can be shown to function in vivo in Drosophila, “humanized” Notch flies may allow functional studies of disease-linked variants/mutations in the context of the human protein in an in vivo environment (Bellen and Yamamoto 2015; Harnish et al. 2019).
Lastly, there are a number of non-canonical Notch signaling events that need further mechanistic studies. For example, Drosophila Notch has been shown to be involved in axon guidance through Abl (Abelson murine leukemia viral oncogene homolog) tyrosine kinase in a manner that does not depend on Su(H) (Giniger 1998; Kannan et al. 2018). In another example in mammalian cells, NICD was found in the mitochondria where it prevented apoptosis in an RBPJ independent fashion (Perumalsamy et al. 2010). How these and other non-canonical functions of Notch are mechanistically achieved and whether such functions are evolutionarily conserved are still open questions that need to be investigated (Alfred and Vaccari 2018). By screening for Notch alleles that exhibit unique phenotypes, uncovering the molecular lesions of such mutations, and further understanding the functional impact of these changes on diverse context of Notch biology, researchers will likely continue to discover unanticipated and exciting aspects of the Notch signaling pathway. The fact that Drosophila is the only genetic model organism that possesses a single gene that encodes a Notch receptor (even C. elegans has two) affords fly geneticists a great advantage. By extrapolating the findings obtained in Drosophila into mammalian systems and combining this knowledge with precise structural information and large-scale human genomics data sets, we will likely identify new biological knowledge that will become useful in understanding and treating genetic disorders or cancers that are linked to defective Notch signaling.
Acknowledgements
I apologize to colleagues whose works I was not able to include in this review article. I thank Dr. Sheng-An Yang, Jose L. Salazar, Samantha L. Deal and J. Michael Harnish for constructive comments and helpful suggestions. S.Y. is currently supported by the following grants from the National Institutes of Health (U54NS093793, R24OD022005, R01DC014932).
References Cited
- Ables JL, Breunig JJ, Eisch AJ, Rakic P (2011) Not(ch) just development: Notch signalling in the adult brain. Nat Rev Neurosci 12 (5):269–283. 10.1038/nrn3024 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Acar M, Jafar-Nejad H, Takeuchi H, Rajan A, Ibrani D, Rana NA, Pan H, Haltiwanger RS, Bellen HJ (2008) Rumi is a CAP10 domain glycosyltransferase that modifies Notch and is required for Notch signaling. Cell 132 (2):247–258 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Agrawal N, Frederick MJ, Pickering CR, Bettegowda C, Chang K, Li RJ, Fakhry C, Xie TX, Zhang J, Wang J, Zhang N, El-Naggar AK, Jasser SA, Weinstein JN, Trevino L, Drummond JA, Muzny DM, Wu Y, Wood LD, Hruban RH, Westra WH, Koch WM, Califano JA, Gibbs RA, Sidransky D, Vogelstein B, Velculescu VE, Papadopoulos N, Wheeler DA, Kinzler KW, Myers JN (2011) Exome sequencing of head and neck squamous cell carcinoma reveals inactivating mutations in NOTCH1. Science 333 (6046):1154–1157 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Akazawa C, Sasai Y, Nakanishi S, Kageyama R (1992) Molecular characterization of a rat negative regulator with a basic helix-loop-helix structure predominantly expressed in the developing nervous system. J Biol Chem 267 (30):21879–21885 [PubMed] [Google Scholar]
- Alfred V, Vaccari T (2018) Mechanisms of Non-canonical Signaling in Health and Disease: Diversity to Take Therapy up a Notch? Adv Exp Med Biol 1066:187–204. 10.1007/978-3-319-89512-3_9 [DOI] [PubMed] [Google Scholar]
- Amberger JS, Bocchini CA, Scott AF, Hamosh A (2019) OMIM.org: leveraging knowledge across phenotype-gene relationships. Nucleic Acids Res 47 (D1):D1038–D1043. 10.1093/nar/gky1151 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arboleda-Velasquez JF, Manent J, Lee JH, Tikka S, Ospina C, Vanderburg CR, Frosch MP, Rodriguez-Falcon M, Villen J, Gygi S, Lopera F, Kalimo H, Moskowitz MA, Ayata C, Louvi A, Artavanis-Tsakonas S (2011) Hypomorphic Notch 3 alleles link Notch signaling to ischemic cerebral small-vessel disease. Proc Natl Acad Sci U S A 108 (21):E128–135. 10.1073/pnas.1101964108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arnett KL, Hass M, McArthur DG, Ilagan MX, Aster JC, Kopan R, Blacklow SC (2010) Structural and mechanistic insights into cooperative assembly of dimeric Notch transcription complexes. Nat Struct Mol Biol 17 (11):1312–1317. 10.1038/nsmb.1938 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Artavanis-Tsakonas S, Muskavitch MA (2010) Notch: the past, the present, and the future. Curr Top Dev Biol 92:1–29 [DOI] [PubMed] [Google Scholar]
- Aster JC, Blacklow SC, Pear WS (2009) Notch signalling in T-cell lymphoblastic leukaemia/lymphoma and other haematological malignancies. J Pathol 223 (2):262–273 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aster JC, Pear WS, Blacklow SC (2017) The Varied Roles of Notch in Cancer. Annu Rev Pathol 12:245–275. 10.1146/annurev-pathol-052016-100127 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aster JC, Simms WB, Zavala-Ruiz Z, Patriub V, North CL, Blacklow SC (1999) The folding and structural integrity of the first LIN-12 module of human Notch1 are calcium-dependent. Biochemistry 38 (15):4736–4742 [DOI] [PubMed] [Google Scholar]
- Axelrod JD, Matsuno K, Artavanis-Tsakonas S, Perrimon N (1996) Interaction between Wingless and Notch signaling pathways mediated by dishevelled. Science 271 (5257):1826–1832. 10.1126/science.271.5257.1826 [DOI] [PubMed] [Google Scholar]
- Bakker H, Oka T, Ashikov A, Yadav A, Berger M, Rana NA, Bai X, Jigami Y, Haltiwanger RS, Esko JD, Gerardy-Schahn R (2009) Functional UDP-xylose transport across the endoplasmic reticulum/Golgi membrane in a Chinese hamster ovary cell mutant defective in UDP-xylose Synthase. J Biol Chem 284 (4):2576–2583 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barolo S, Stone T, Bang AG, Posakony JW (2002) Default repression and Notch signaling: Hairless acts as an adaptor to recruit the corepressors Groucho and dCtBP to Suppressor of Hairless. Genes Dev 16 (15):1964–1976 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baron M (2017) Combining genetic and biophysical approaches to probe the structure and function relationships of the notch receptor. Mol Membr Biol 34 (1–2):33–49. 10.1080/09687688.2018.1503742 [DOI] [PubMed] [Google Scholar]
- Baudrimont M, Dubas F, Joutel A, Tournier-Lasserve E, Bousser MG (1993) Autosomal dominant leukoencephalopathy and subcortical ischemic stroke. A clinicopathological study. Stroke 24 (1):122–125 [DOI] [PubMed] [Google Scholar]
- Becam I, Fiuza UM, Arias AM, Milan M (2010) A role of receptor Notch in ligand cis-inhibition in Drosophila. Curr Biol 20 (6):554–560 [DOI] [PubMed] [Google Scholar]
- Bellen HJ, Wangler MF, Yamamoto S (2019) The fruit fly at the interface of diagnosis and pathogenic mechanisms of rare and common human diseases. Hum Mol Genet. 10.1093/hmg/ddz135 [DOI] [PMC free article] [PubMed]
- Bellen HJ, Yamamoto S (2015) Morgan’s legacy: fruit flies and the functional annotation of conserved genes. Cell 163 (1):12–14. 10.1016/j.cell.2015.09.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bier E (2005) Drosophila, the golden bug, emerges as a tool for human genetics. Nat Rev Genet 6 (1):9–23. 10.1038/nrg1503 [DOI] [PubMed] [Google Scholar]
- Bigas A, Porcheri C (2018) Notch and Stem Cells. Adv Exp Med Biol 1066:235–263. 10.1007/978-3-319-89512-3_12 [DOI] [PubMed] [Google Scholar]
- Bishop SA, Klein T, Arias AM, Couso JP (1999) Composite signalling from Serrate and Delta establishes leg segments in Drosophila through Notch. Development (Cambridge, England) 126 (13):2993–3003 [DOI] [PubMed] [Google Scholar]
- Blaumueller CM, Qi H, Zagouras P, Artavanis-Tsakonas S (1997) Intracellular cleavage of Notch leads to a heterodimeric receptor on the plasma membrane. Cell 90 (2):281–291 [DOI] [PubMed] [Google Scholar]
- Borggrefe T, Lauth M, Zwijsen A, Huylebroeck D, Oswald F, Giaimo BD (2016) The Notch intracellular domain integrates signals from Wnt, Hedgehog, TGFbeta/BMP and hypoxia pathways. Biochim Biophys Acta 1863 (2):303–313. 10.1016/j.bbamcr.2015.11.020 [DOI] [PubMed] [Google Scholar]
- Bray SJ (2006) Notch signalling: a simple pathway becomes complex. Nat Rev Mol Cell Biol 7 (9):678–689 [DOI] [PubMed] [Google Scholar]
- Bray SJ, Gomez-Lamarca M (2018) Notch after cleavage. Curr Opin Cell Biol 51:103–109. 10.1016/j.ceb.2017.12.008 [DOI] [PubMed] [Google Scholar]
- Brennan K, Gardner P (2002) Notching up another pathway. Bioessays 24 (5):405–410 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brennan K, Tateson R, Lewis K, Arias AM (1997) A functional analysis of Notch mutations in Drosophila. Genetics 147 (1):177–188 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brennan K, Tateson R, Lieber T, Couso JP, Zecchini V, Arias AM (1999) The abruptex mutations of notch disrupt the establishment of proneural clusters in Drosophila. Dev Biol 216 (1):230–242. 10.1006/dbio.1999.9501 [DOI] [PubMed] [Google Scholar]
- Bruckner K, Perez L, Clausen H, Cohen S (2000) Glycosyltransferase activity of Fringe modulates Notch-Delta interactions. Nature 406 (6794):411–415 [DOI] [PubMed] [Google Scholar]
- Cagan RL, Ready DF (1989) Notch is required for successive cell decisions in the developing Drosophila retina. Genes Dev 3 (8):1099–1112. 10.1101/gad.3.8.1099 [DOI] [PubMed] [Google Scholar]
- Cancer Genome Atlas Research N, Brat DJ, Verhaak RG, Aldape KD, Yung WK, Salama SR, Cooper LA, Rheinbay E, Miller CR, Vitucci M, Morozova O, Robertson AG, Noushmehr H, Laird PW, Cherniack AD, Akbani R, Huse JT, Ciriello G, Poisson LM, Barnholtz-Sloan JS, Berger MS, Brennan C, Colen RR, Colman H, Flanders AE, Giannini C, Grifford M, Iavarone A, Jain R, Joseph I, Kim J, Kasaian K, Mikkelsen T, Murray BA, O’Neill BP, Pachter L, Parsons DW, Sougnez C, Sulman EP, Vandenberg SR, Van Meir EG, von Deimling A, Zhang H, Crain D, Lau K, Mallery D, Morris S, Paulauskis J, Penny R, Shelton T, Sherman M, Yena P, Black A, Bowen J, Dicostanzo K, Gastier-Foster J, Leraas KM, Lichtenberg TM, Pierson CR, Ramirez NC, Taylor C, Weaver S, Wise L, Zmuda E, Davidsen T, Demchok JA, Eley G, Ferguson ML, Hutter CM, Mills Shaw KR, Ozenberger BA, Sheth M, Sofia HJ, Tarnuzzer R, Wang Z, Yang L, Zenklusen JC, Ayala B, Baboud J, Chudamani S, Jensen MA, Liu J, Pihl T, Raman R, Wan Y, Wu Y, Ally A, Auman JT, Balasundaram M, Balu S, Baylin SB, Beroukhim R, Bootwalla MS, Bowlby R, Bristow CA, Brooks D, Butterfield Y, Carlsen R, Carter S, Chin L, Chu A, Chuah E, Cibulskis K, Clarke A, Coetzee SG, Dhalla N, Fennell T, Fisher S, Gabriel S, Getz G, Gibbs R, Guin R, Hadjipanayis A, Hayes DN, Hinoue T, Hoadley K, Holt RA, Hoyle AP, Jefferys SR, Jones S, Jones CD, Kucherlapati R, Lai PH, Lander E, Lee S, Lichtenstein L, Ma Y, Maglinte DT, Mahadeshwar HS, Marra MA, Mayo M, Meng S, Meyerson ML, Mieczkowski PA, Moore RA, Mose LE, Mungall AJ, Pantazi A, Parfenov M, Park PJ, Parker JS, Perou CM, Protopopov A, Ren X, Roach J, Sabedot TS, Schein J, Schumacher SE, Seidman JG, Seth S, Shen H, Simons JV, Sipahimalani P, Soloway MG, Song X, Sun H, Tabak B, Tam A, Tan D, Tang J, Thiessen N, Triche T Jr., Van Den Berg DJ, Veluvolu U, Waring S, Weisenberger DJ, Wilkerson MD, Wong T, Wu J, Xi L, Xu AW, Yang L, Zack TI, Zhang J, Aksoy BA, Arachchi H, Benz C, Bernard B, Carlin D, Cho J, DiCara D, Frazer S, Fuller GN, Gao J, Gehlenborg N, Haussler D, Heiman DI, Iype L, Jacobsen A, Ju Z, Katzman S, Kim H, Knijnenburg T, Kreisberg RB, Lawrence MS, Lee W, Leinonen K, Lin P, Ling S, Liu W, Liu Y, Liu Y, Lu Y, Mills G, Ng S, Noble MS, Paull E, Rao A, Reynolds S, Saksena G, Sanborn Z, Sander C, Schultz N, Senbabaoglu Y, Shen R, Shmulevich I, Sinha R, Stuart J, Sumer SO, Sun Y, Tasman N, Taylor BS, Voet D, Weinhold N, Weinstein JN, Yang D, Yoshihara K, Zheng S, Zhang W, Zou L, Abel T, Sadeghi S, Cohen ML, Eschbacher J, Hattab EM, Raghunathan A, Schniederjan MJ, Aziz D, Barnett G, Barrett W, Bigner DD, Boice L, Brewer C, Calatozzolo C, Campos B, Carlotti CG Jr., Chan TA, Cuppini L, Curley E, Cuzzubbo S, Devine K, DiMeco F, Duell R, Elder JB, Fehrenbach A, Finocchiaro G, Friedman W, Fulop J, Gardner J, Hermes B, Herold-Mende C, Jungk C, Kendler A, Lehman NL, Lipp E, Liu O, Mandt R, McGraw M, McLendon R, McPherson C, Neder L, Nguyen P, Noss A, Nunziata R, Ostrom QT, Palmer C, Perin A, Pollo B, Potapov A, Potapova O, Rathmell WK, Rotin D, Scarpace L, Schilero C, Senecal K, Shimmel K, Shurkhay V, Sifri S, Singh R, Sloan AE, Smolenski K, Staugaitis SM, Steele R, Thorne L, Tirapelli DP, Unterberg A, Vallurupalli M, Wang Y, Warnick R, Williams F, Wolinsky Y, Bell S, Rosenberg M, Stewart C, Huang F, Grimsby JL, Radenbaugh AJ, Zhang J (2015) Comprehensive, Integrative Genomic Analysis of Diffuse Lower-Grade Gliomas. N Engl J Med 372 (26):2481–2498. 10.1056/NEJMoa1402121 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cardinale CJ, Li D, Tian L, Connolly JJ, March ME, Hou C, Wang F, Snyder J, Kim CE, Chiavacci RM, Sleiman PM, Burnham JM, Hakonarson H (2016) Association of a rare NOTCH4 coding variant with systemic sclerosis: a family-based whole exome sequencing study. BMC Musculoskelet Disord 17 (1):462 10.1186/s12891-016-1320-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Charng WL, Yamamoto S, Bellen HJ (2014a) Shared mechanisms between Drosophila peripheral nervous system development and human neurodegenerative diseases. Curr Opin Neurobiol 27:158–164. 10.1016/j.conb.2014.03.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Charng WL, Yamamoto S, Jaiswal M, Bayat V, Xiong B, Zhang K, Sandoval H, David G, Gibbs S, Lu HC, Chen K, Giagtzoglou N, Bellen HJ (2014b) Drosophila Tempura, a novel protein prenyltransferase alpha subunit, regulates notch signaling via Rab1 and Rab11. PLoS Biol 12 (1):e1001777 10.1371/journal.pbio.1001777 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chaturvedi D, Reichert H, Gunage RD, VijayRaghavan K (2017) Identification and functional characterization of muscle satellite cells in Drosophila. Elife 6 10.7554/eLife.30107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Collu GM, Hidalgo-Sastre A, Brennan K (2014) Wnt-Notch signalling crosstalk in development and disease. Cell Mol Life Sci 71 (18):3553–3567. 10.1007/s00018-014-1644-x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cordle J, Johnson S, Tay JZ, Roversi P, Wilkin MB, de Madrid BH, Shimizu H, Jensen S, Whiteman P, Jin B, Redfield C, Baron M, Lea SM, Handford PA (2008) A conserved face of the Jagged/Serrate DSL domain is involved in Notch trans-activation and cis-inhibition. Nat Struct Mol Biol 15 (8):849–857 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Couso JP, Martinez Arias A (1994) Notch is required for wingless signaling in the epidermis of Drosophila. Cell 79 (2):259–272 [DOI] [PubMed] [Google Scholar]
- D’Souza B, Meloty-Kapella L, Weinmaster G (2010) Canonical and non-canonical Notch ligands. Curr Top Dev Biol 92:73–129 [DOI] [PMC free article] [PubMed] [Google Scholar]
- de Celis JF, Barrio R, del Arco A, Garcia-Bellido A (1993) Genetic and molecular characterization of a Notch mutation in its Delta- and Serrate-binding domain in Drosophila. Proc Natl Acad Sci U S A 90 (9):4037–4041 [DOI] [PMC free article] [PubMed] [Google Scholar]
- de Celis JF, Bray SJ (2000) The Abruptex domain of Notch regulates negative interactions between Notch, its ligands and Fringe. Development (Cambridge, England) 127 (6):1291–1302 [DOI] [PubMed] [Google Scholar]
- de Celis JF, Garcia-Bellido A (1994) Modifications of the notch function by Abruptex mutations in Drosophila melanogaster. Genetics 136 (1):183–194 [DOI] [PMC free article] [PubMed] [Google Scholar]
- De Strooper B, Annaert W, Cupers P, Saftig P, Craessaerts K, Mumm JS, Schroeter EH, Schrijvers V, Wolfe MS, Ray WJ, Goate A, Kopan R (1999) A presenilin-1-dependent gamma-secretase-like protease mediates release of Notch intracellular domain. Nature 398 (6727):518–522 [DOI] [PubMed] [Google Scholar]
- Deal SL, Yamamoto S (2018) Unraveling Novel Mechanisms of Neurodegeneration Through a Large-Scale Forward Genetic Screen in Drosophila. Front Genet 9:700 10.3389/fgene.2018.00700 [DOI] [PMC free article] [PubMed] [Google Scholar]
- del Alamo D, Rouault H, Schweisguth F (2011) Mechanism and significance of cis-inhibition in Notch signalling. Curr Biol 21 (1):R40–47 [DOI] [PubMed] [Google Scholar]
- Dexter JS (1914) The Analysis of a Case of Continuous Variation in Drosophila by a Study of Its Linkage Relations. The American Naturalist 48 (576):712–758 [Google Scholar]
- Diederich RJ, Matsuno K, Hing H, Artavanis-Tsakonas S (1994) Cytosolic interaction between deltex and Notch ankyrin repeats implicates deltex in the Notch signaling pathway. Development (Cambridge, England) 120 (3):473–481 [DOI] [PubMed] [Google Scholar]
- Espinosa L, Ingles-Esteve J, Aguilera C, Bigas A (2003) Phosphorylation by glycogen synthase kinase-3 beta down-regulates Notch activity, a link for Notch and Wnt pathways. J Biol Chem 278 (34):32227–32235. 10.1074/jbc.M304001200 [DOI] [PubMed] [Google Scholar]
- Fabbri G, Rasi S, Rossi D, Trifonov V, Khiabanian H, Ma J, Grunn A, Fangazio M, Capello D, Monti S, Cresta S, Gargiulo E, Forconi F, Guarini A, Arcaini L, Paulli M, Laurenti L, Larocca LM, Marasca R, Gattei V, Oscier D, Bertoni F, Mullighan CG, Foa R, Pasqualucci L, Rabadan R, Dalla-Favera R, Gaidano G (2011) Analysis of the chronic lymphocytic leukemia coding genome: role of NOTCH1 mutational activation. J Exp Med 208 (7):1389–1401 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fernandez-Valdivia R, Takeuchi H, Samarghandi A, Lopez M, Leonardi J, Haltiwanger RS, Jafar-Nejad H (2011) Regulation of mammalian Notch signaling and embryonic development by the protein O-glucosyltransferase Rumi. Development (Cambridge, England) 138 (10):1925–1934 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fleming RJ, Gu Y, Hukriede NA (1997) Serrate-mediated activation of Notch is specifically blocked by the product of the gene fringe in the dorsal compartment of the Drosophila wing imaginal disc. Development (Cambridge, England) 124 (15):2973–2981 [DOI] [PubMed] [Google Scholar]
- Fleming RJ, Scottgale TN, Diederich RJ, Artavanis-Tsakonas S (1990) The gene Serrate encodes a putative EGF-like transmembrane protein essential for proper ectodermal development in Drosophila melanogaster. Genes Dev 4 (12A):2188–2201. 10.1101/gad.4.12a.2188 [DOI] [PubMed] [Google Scholar]
- Foster GG (1975) Negative complementation at the notch locus of Drosophila melanogaster. Genetics 81 (1):99–120 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fryer CJ, Lamar E, Turbachova I, Kintner C, Jones KA (2002) Mastermind mediates chromatin-specific transcription and turnover of the Notch enhancer complex. Genes Dev 16 (11):1397–1411 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fryer CJ, White JB, Jones KA (2004) Mastermind recruits CycC:CDK8 to phosphorylate the Notch ICD and coordinate activation with turnover. Mol Cell 16 (4):509–520 [DOI] [PubMed] [Google Scholar]
- Fuwa TJ, Hori K, Sasamura T, Higgs J, Baron M, Matsuno K (2006) The first deltex null mutant indicates tissue-specific deltex-dependent Notch signaling in Drosophila. Mol Genet Genomics 275 (3):251–263 [DOI] [PubMed] [Google Scholar]
- Garg V, Muth AN, Ransom JF, Schluterman MK, Barnes R, King IN, Grossfeld PD, Srivastava D (2005) Mutations in NOTCH1 cause aortic valve disease. Nature 437 (7056):270–274. 10.1038/nature03940 [DOI] [PubMed] [Google Scholar]
- Ge C, Liu T, Hou X, Stanley P (2008) In vivo consequences of deleting EGF repeats 8–12 including the ligand binding domain of mouse Notch1. BMC Dev Biol 8:48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- George J, Lim JS, Jang SJ, Cun Y, Ozretic L, Kong G, Leenders F, Lu X, Fernandez-Cuesta L, Bosco G, Muller C, Dahmen I, Jahchan NS, Park KS, Yang D, Karnezis AN, Vaka D, Torres A, Wang MS, Korbel JO, Menon R, Chun SM, Kim D, Wilkerson M, Hayes N, Engelmann D, Putzer B, Bos M, Michels S, Vlasic I, Seidel D, Pinther B, Schaub P, Becker C, Altmuller J, Yokota J, Kohno T, Iwakawa R, Tsuta K, Noguchi M, Muley T, Hoffmann H, Schnabel PA, Petersen I, Chen Y, Soltermann A, Tischler V, Choi CM, Kim YH, Massion PP, Zou Y, Jovanovic D, Kontic M, Wright GM, Russell PA, Solomon B, Koch I, Lindner M, Muscarella LA, la Torre A, Field JK, Jakopovic M, Knezevic J, Castanos-Velez E, Roz L, Pastorino U, Brustugun OT, Lund-Iversen M, Thunnissen E, Kohler J, Schuler M, Botling J, Sandelin M, Sanchez-Cespedes M, Salvesen HB, Achter V, Lang U, Bogus M, Schneider PM, Zander T, Ansen S, Hallek M, Wolf J, Vingron M, Yatabe Y, Travis WD, Nurnberg P, Reinhardt C, Perner S, Heukamp L, Buttner R, Haas SA, Brambilla E, Peifer M, Sage J, Thomas RK (2015) Comprehensive genomic profiles of small cell lung cancer. Nature 524 (7563):47–53. 10.1038/nature14664 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giniger E (1998) A role for Abl in Notch signaling. Neuron 20 (4):667–681 [DOI] [PubMed] [Google Scholar]
- Gordon WR, Vardar-Ulu D, Histen G, Sanchez-Irizarry C, Aster JC, Blacklow SC (2007) Structural basis for autoinhibition of Notch. Nat Struct Mol Biol 14 (4):295–300 [DOI] [PubMed] [Google Scholar]
- Gottschewski G (1935) New mutants report. Drosophila Information Service 4:14–16 [Google Scholar]
- Gripp KW, Robbins KM, Sobreira NL, Witmer PD, Bird LM, Avela K, Makitie O, Alves D, Hogue JS, Zackai EH, Doheny KF, Stabley DL, Sol-Church K (2015) Truncating mutations in the last exon of NOTCH3 cause lateral meningocele syndrome. Am J Med Genet A 167A (2):271–281. 10.1002/ajmg.a.36863 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guarani V, Deflorian G, Franco CA, Kruger M, Phng LK, Bentley K, Toussaint L, Dequiedt F, Mostoslavsky R, Schmidt MHH, Zimmermann B, Brandes RP, Mione M, Westphal CH, Braun T, Zeiher AM, Gerhardt H, Dimmeler S, Potente M (2011) Acetylation-dependent regulation of endothelial Notch signalling by the SIRT1 deacetylase. Nature 473 (7346):234–238. 10.1038/nature09917 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gupta-Rossi N, Le Bail O, Gonen H, Brou C, Logeat F, Six E, Ciechanover A, Israel A (2001) Functional interaction between SEL-10, an F-box protein, and the nuclear form of activated Notch1 receptor. J Biol Chem 276 (37):34371–34378 [DOI] [PubMed] [Google Scholar]
- Guruharsha KG, Kankel MW, Artavanis-Tsakonas S (2012) The Notch signalling system: recent insights into the complexity of a conserved pathway. Nat Rev Genet 13 (9):654–666. 10.1038/nrg3272 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hambleton S, Valeyev NV, Muranyi A, Knott V, Werner JM, McMichael AJ, Handford PA, Downing AK (2004) Structural and functional properties of the human notch-1 ligand binding region. Structure 12 (12):2173–2183 [DOI] [PubMed] [Google Scholar]
- Handford PA, Korona B, Suckling R, Redfield C, Lea SM (2018) Structural Insights into Notch Receptor-Ligand Interactions. Adv Exp Med Biol 1066:33–46. 10.1007/978-3-319-89512-3_2 [DOI] [PubMed] [Google Scholar]
- Handford PA, Mayhew M, Baron M, Winship PR, Campbell ID, Brownlee GG (1991) Key residues involved in calcium-binding motifs in EGF-like domains. Nature 351 (6322):164–167. 10.1038/351164a0 [DOI] [PubMed] [Google Scholar]
- Hara K, Shiga A, Fukutake T, Nozaki H, Miyashita A, Yokoseki A, Kawata H, Koyama A, Arima K, Takahashi T, Ikeda M, Shiota H, Tamura M, Shimoe Y, Hirayama M, Arisato T, Yanagawa S, Tanaka A, Nakano I, Ikeda S, Yoshida Y, Yamamoto T, Ikeuchi T, Kuwano R, Nishizawa M, Tsuji S, Onodera O (2009) Association of HTRA1 mutations and familial ischemic cerebral small-vessel disease. N Engl J Med 360 (17):1729–1739. 10.1056/NEJMoa0801560 [DOI] [PubMed] [Google Scholar]
- Harnish JM, Deal SL, Chao HT, Wangler MF, Yamamoto S (2019) In vivo functional study of disease-associated rare human variants using Drosophila. J Vis Exp In Press [DOI] [PMC free article] [PubMed]
- Hartenstein V, Posakony JW (1990) A dual function of the Notch gene in Drosophila sensillum development. Dev Biol 142 (1):13–30 [DOI] [PubMed] [Google Scholar]
- Hartley DA, Xu TA, Artavanis-Tsakonas S (1987) The embryonic expression of the Notch locus of Drosophila melanogaster and the implications of point mutations in the extracellular EGF-like domain of the predicted protein. EMBO J 6 (11):3407–3417 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Harvey BM, Haltiwanger RS (2018) Regulation of Notch Function by O-Glycosylation. Adv Exp Med Biol 1066:59–78. 10.1007/978-3-319-89512-3_4 [DOI] [PubMed] [Google Scholar]
- Harvey BM, Rana NA, Moss H, Leonardi J, Jafar-Nejad H, Haltiwanger RS (2016) Mapping Sites of O-Glycosylation and Fringe Elongation on Drosophila Notch. J Biol Chem 291 (31):16348–16360. 10.1074/jbc.M116.732537 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hein K, Mittler G, Cizelsky W, Kuhl M, Ferrante F, Liefke R, Berger IM, Just S, Strang JE, Kestler HA, Oswald F, Borggrefe T (2015) Site-specific methylation of Notch1 controls the amplitude and duration of the Notch1 response. Sci Signal 8 (369): 10.1126/scisignal.2005892 [DOI] [PubMed] [Google Scholar]
- Heitzler P (2010) Biodiversity and noncanonical Notch signaling. Curr Top Dev Biol 92:457–481. 10.1016/S0070-2153(10)92014-0 [DOI] [PubMed] [Google Scholar]
- Heitzler P, Simpson P (1993) Altered epidermal growth factor-like sequences provide evidence for a role of Notch as a receptor in cell fate decisions. Development (Cambridge, England) 117 (3):1113–1123 [DOI] [PubMed] [Google Scholar]
- Hing HK, Bangalore L, Sun X, Artavanis-Tsakonas S (1999) Mutations in the heatshock cognate 70 protein (hsc4) modulate Notch signaling. Eur J Cell Biol 78 (10):690–697. 10.1016/S0171-9335(99)80037-5 [DOI] [PubMed] [Google Scholar]
- Hoppe PE, Greenspan RJ (1986) Local function of the Notch gene for embryonic ectodermal pathway choice in Drosophila. Cell 46 (5):773–783 [DOI] [PubMed] [Google Scholar]
- Huenniger K, Kramer A, Soom M, Chang I, Kohler M, Depping R, Kehlenbach RH, Kaether C (2010) Notch1 signaling is mediated by importins alpha 3, 4, and 7. Cell Mol Life Sci 67 (18):3187–3196 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Irvine KD, Wieschaus E (1994) fringe, a Boundary-specific signaling molecule, mediates interactions between dorsal and ventral cells during Drosophila wing development. Cell 79 (4):595–606 [DOI] [PubMed] [Google Scholar]
- Ishiko A, Shimizu A, Nagata E, Takahashi K, Tabira T, Suzuki N (2006) Notch3 ectodomain is a major component of granular osmiophilic material (GOM) in CADASIL. Acta Neuropathol 112 (3):333–339. 10.1007/s00401-006-0116-2 [DOI] [PubMed] [Google Scholar]
- Ishio A, Sasamura T, Ayukawa T, Kuroda J, Ishikawa HO, Aoyama N, Matsumoto K, Gushiken T, Okajima T, Yamakawa T, Matsuno K (2015) O-fucose monosaccharide of Drosophila Notch has a temperature-sensitive function and cooperates with O-glucose glycan in Notch transport and Notch signaling activation. J Biol Chem 290 (1):505–519. 10.1074/jbc.M114.616847 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Isidor B, Lindenbaum P, Pichon O, Bezieau S, Dina C, Jacquemont S, Martin-Coignard D, Thauvin-Robinet C, Le Merrer M, Mandel JL, David A, Faivre L, Cormier-Daire V, Redon R, Le Caignec C (2011) Truncating mutations in the last exon of NOTCH2 cause a rare skeletal disorder with osteoporosis. Nat Genet 43 (4):306–308. 10.1038/ng.778 [DOI] [PubMed] [Google Scholar]
- Jafar-Nejad H, Leonardi J, Fernandez-Valdivia R (2010) Role of glycans and glycosyltransferases in the regulation of Notch signaling. Glycobiology 20 (8):931–949 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jorissen E, De Strooper B (2010) Gamma-secretase and the intramembrane proteolysis of Notch. Curr Top Dev Biol 92:201–230 [DOI] [PubMed] [Google Scholar]
- Joutel A, Corpechot C, Ducros A, Vahedi K, Chabriat H, Mouton P, Alamowitch S, Domenga V, Cecillion M, Marechal E, Maciazek J, Vayssiere C, Cruaud C, Cabanis EA, Ruchoux MM, Weissenbach J, Bach JF, Bousser MG, Tournier-Lasserve E (1996) Notch3 mutations in CADASIL, a hereditary adult-onset condition causing stroke and dementia. Nature 383 (6602):707–710 [DOI] [PubMed] [Google Scholar]
- Joutel A, Vahedi K, Corpechot C, Troesch A, Chabriat H, Vayssiere C, Cruaud C, Maciazek J, Weissenbach J, Bousser MG, Bach JF, Tournier-Lasserve E (1997) Strong clustering and stereotyped nature of Notch3 mutations in CADASIL patients. Lancet 350 (9090):1511–1515. 10.1016/S0140-6736(97)08083-5 [DOI] [PubMed] [Google Scholar]
- Kannan R, Cox E, Wang L, Kuzina I, Gu Q, Giniger E (2018) Tyrosine phosphorylation and proteolytic cleavage of Notch are required for non-canonical Notch/Abl signaling in Drosophila axon guidance. Development (Cambridge, England) 145 (2). 10.1242/dev.151548 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kao HY, Ordentlich P, Koyano-Nakagawa N, Tang Z, Downes M, Kintner CR, Evans RM, Kadesch T (1998) A histone deacetylase corepressor complex regulates the Notch signal transduction pathway. Genes Dev 12 (15):2269–2277 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kelley MR, Kidd S, Deutsch WA, Young MW (1987) Mutations altering the structure of epidermal growth factor-like coding sequences at the Drosophila Notch locus. Cell 51 (4):539–548 [DOI] [PubMed] [Google Scholar]
- Kelly DF, Lake RJ, Middelkoop TC, Fan HY, Artavanis-Tsakonas S, Walz T (2010) Molecular structure and dimeric organization of the Notch extracellular domain as revealed by electron microscopy. PloS one 5 (5):e10532 10.1371/journal.pone.0010532 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kidd S, Baylies MK, Gasic GP, Young MW (1989) Structure and distribution of the Notch protein in developing Drosophila. Genes Dev 3 (8):1113–1129. 10.1101/gad.3.8.1113 [DOI] [PubMed] [Google Scholar]
- Kidd S, Lockett TJ, Young MW (1983) The Notch locus of Drosophila melanogaster. Cell 34 (2):421–433 [DOI] [PubMed] [Google Scholar]
- Kiel MJ, Velusamy T, Betz BL, Zhao L, Weigelin HG, Chiang MY, Huebner-Chan DR, Bailey NG, Yang DT, Bhagat G, Miranda RN, Bahler DW, Medeiros LJ, Lim MS, Elenitoba-Johnson KS (2012) Whole-genome sequencing identifies recurrent somatic NOTCH2 mutations in splenic marginal zone lymphoma. J Exp Med 209 (9):1553–1565. 10.1084/jem.20120910 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Klinakis A, Lobry C, Abdel-Wahab O, Oh P, Haeno H, Buonamici S, van De Walle I, Cathelin S, Trimarchi T, Araldi E, Liu C, Ibrahim S, Beran M, Zavadil J, Efstratiadis A, Taghon T, Michor F, Levine RL, Aifantis I (2011) A novel tumour-suppressor function for the Notch pathway in myeloid leukaemia. Nature 473 (7346):230–233. 10.1038/nature09999 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Knott GJ, Doudna JA (2018) CRISPR-Cas guides the future of genetic engineering. Science 361 (6405):866–869. 10.1126/science.aat5011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koenig SN, LaHaye S, Feller JD, Rowland P, Hor KN, Trask AJ, Janssen PM, Radtke F, Lilly B, Garg V (2017) Notch1 haploinsufficiency causes ascending aortic aneurysms in mice. JCI Insight 2 (21). 10.1172/jci.insight.91353 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kopan R (1999) All good things must come to an end: how is Notch signaling turned off? Sci STKE 1999 (9): 10.1126/stke.1999.9.pe1 [DOI] [PubMed] [Google Scholar]
- Kopan R, Ilagan MX (2009) The canonical Notch signaling pathway: unfolding the activation mechanism. Cell 137 (2):216–233 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kopan R, Nye JS, Weintraub H (1994) The intracellular domain of mouse Notch: a constitutively activated repressor of myogenesis directed at the basic helix-loop-helix region of MyoD. Development (Cambridge, England) 120 (9):2385–2396 [DOI] [PubMed] [Google Scholar]
- Kopczynski CC, Alton AK, Fechtel K, Kooh PJ, Muskavitch MA (1988) Delta, a Drosophila neurogenic gene, is transcriptionally complex and encodes a protein related to blood coagulation factors and epidermal growth factor of vertebrates. Genes & Development 2 (12b):1723–1735 [DOI] [PubMed] [Google Scholar]
- Kovall RA, Blacklow SC (2010) Mechanistic insights into Notch receptor signaling from structural and biochemical studies. Curr Top Dev Biol 92:31–71 [DOI] [PubMed] [Google Scholar]
- Kovall RA, Hendrickson WA (2004) Crystal structure of the nuclear effector of Notch signaling, CSL, bound to DNA. Embo J 23 (17):3441–3451 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kurooka H, Honjo T (2000) Functional interaction between the mouse notch1 intracellular region and histone acetyltransferases PCAF and GCN5. J Biol Chem 275 (22):17211–17220 [DOI] [PubMed] [Google Scholar]
- Lake RJ, Grimm LM, Veraksa A, Banos A, Artavanis-Tsakonas S (2009) In vivo analysis of the Notch receptor S1 cleavage. PloS one 4 (8):e6728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee TV, Pandey A, Jafar-Nejad H (2017) Xylosylation of the Notch receptor preserves the balance between its activation by trans-Delta and inhibition by cis-ligands in Drosophila. PLoS Genet 13 (4):e1006723 10.1371/journal.pgen.1006723 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee TV, Sethi MK, Leonardi J, Rana NA, Buettner FF, Haltiwanger RS, Bakker H, Jafar-Nejad H (2013) Negative regulation of notch signaling by xylose. PLoS Genet 9 (6):e1003547 10.1371/journal.pgen.1003547 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lees AD, Waddington CH (1942) The development of the bristles in normal and some mutant types of Drosophila melanogaster. Proc R Soc Lond B 131:87–110 [Google Scholar]
- Lehmann R, Dietrich U, Jiménez F, Campos-Ortega JA (1981) Mutations of early neurogenesis inDrosophila. Wilhelm Roux’s archives of developmental biology 190 (4):226–229 [DOI] [PubMed] [Google Scholar]
- Lehmann R, Jimenez F, Dietrich U, Campos-Ortega JA (1983) On the phenotype and development of mutants of early neurogenesis inDrosophila melanogaster. Wilehm Roux Arch Dev Biol 192 (2):62–74. 10.1007/BF00848482 [DOI] [PubMed] [Google Scholar]
- Leimeister C, Externbrink A, Klamt B, Gessler M (1999) Hey genes: a novel subfamily of hairy- and Enhancer of split related genes specifically expressed during mouse embryogenesis. Mech Dev 85 (1–2):173–177 [DOI] [PubMed] [Google Scholar]
- Leonardi J, Fernandez-Valdivia R, Li YD, Simcox AA, Jafar-Nejad H (2011) Multiple O-glucosylation sites on Notch function as a buffer against temperature-dependent loss of signaling. Development (Cambridge, England) 138 (16):3569–3578 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Letai A (2017) Functional precision cancer medicine-moving beyond pure genomics. Nat Med 23 (9):1028–1035. 10.1038/nm.4389 [DOI] [PubMed] [Google Scholar]
- Levitan D, Greenwald I (1995) Facilitation of lin-12-mediated signalling by sel-12, a Caenorhabditis elegans S182 Alzheimer’s disease gene. Nature 377 (6547):351–354 [DOI] [PubMed] [Google Scholar]
- Li X, Cassidy JJ, Reinke CA, Fischboeck S, Carthew RW (2009) A microRNA imparts robustness against environmental fluctuation during development. Cell 137 (2):273–282. 10.1016/j.cell.2009.01.058 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lindsley DL, Zimm GG (1992) The Genome of Drosophila melanogaster. Academic Press, [Google Scholar]
- Logeat F, Bessia C, Brou C, LeBail O, Jarriault S, Seidah NG, Israel A (1998) The Notch1 receptor is cleaved constitutively by a furin-like convertase. Proc Natl Acad Sci U S A 95 (14):8108–8112 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Louvi A, Artavanis-Tsakonas S (2012) Notch and disease: a growing field. Semin Cell Dev Biol 23 (4):473–480. 10.1016/j.semcdb.2012.02.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lovendahl KN, Blacklow SC, Gordon WR (2018) The Molecular Mechanism of Notch Activation. Adv Exp Med Biol 1066:47–58. 10.1007/978-3-319-89512-3_3 [DOI] [PubMed] [Google Scholar]
- Luca VC, Jude KM, Pierce NW, Nachury MV, Fischer S, Garcia KC (2015) Structural basis for Notch1 engagement of Delta-like 4. Science 347 (6224):847–853. 10.1126/science.1261093 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luca VC, Kim BC, Ge C, Kakuda S, Wu D, Roein-Peikar M, Haltiwanger RS, Zhu C, Ha T, Garcia KC (2017) Notch-Jagged complex structure implicates a catch bond in tuning ligand sensitivity. Science 355 (6331):1320–1324. 10.1126/science.aaf9739 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lyman D, Young MW (1993) Further evidence for function of the Drosophila Notch protein as a transmembrane receptor. Proc Natl Acad Sci U S A 90 (21):10395–10399. 10.1073/pnas.90.21.10395 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martignetti JA, Tian L, Li D, Ramirez MC, Camacho-Vanegas O, Camacho SC, Guo Y, Zand DJ, Bernstein AM, Masur SK, Kim CE, Otieno FG, Hou C, Abdel-Magid N, Tweddale B, Metry D, Fournet JC, Papp E, McPherson EW, Zabel C, Vaksmann G, Morisot C, Keating B, Sleiman PM, Cleveland JA, Everman DB, Zackai E, Hakonarson H (2013) Mutations in PDGFRB cause autosomal-dominant infantile myofibromatosis. Am J Hum Genet 92 (6):1001–1007. 10.1016/j.ajhg.2013.04.024 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Masek J, Andersson ER (2017) The developmental biology of genetic Notch disorders. Development (Cambridge, England) 144 (10):1743–1763. 10.1242/dev.148007 [DOI] [PubMed] [Google Scholar]
- Matsumoto K, Ayukawa T, Ishio A, Sasamura T, Yamakawa T, Matsuno K (2016) Dual Roles of O-Glucose Glycans Redundant with Monosaccharide O-Fucose on Notch in Notch Trafficking. J Biol Chem 291 (26):13743–13752. 10.1074/jbc.M115.710483 [DOI] [PMC free article] [PubMed] [Google Scholar]
- McCarter AC, Wang Q, Chiang M (2018) Notch in Leukemia. Adv Exp Med Biol 1066:355–394. 10.1007/978-3-319-89512-3_18 [DOI] [PubMed] [Google Scholar]
- McDaniell R, Warthen DM, Sanchez-Lara PA, Pai A, Krantz ID, Piccoli DA, Spinner NB (2006) NOTCH2 mutations cause Alagille syndrome, a heterogeneous disorder of the notch signaling pathway. Am J Hum Genet 79 (1):169–173 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Micchelli CA, Blair SS (1999) Dorsoventral lineage restriction in wing imaginal discs requires Notch. Nature 401 (6752):473–476. 10.1038/46779 [DOI] [PubMed] [Google Scholar]
- Micchelli CA, Perrimon N (2006) Evidence that stem cells reside in the adult Drosophila midgut epithelium. Nature 439 (7075):475–479. 10.1038/nature04371 [DOI] [PubMed] [Google Scholar]
- Micchelli CA, Rulifson EJ, Blair SS (1997) The function and regulation of cut expression on the wing margin of Drosophila: Notch, Wingless and a dominant negative role for Delta and Serrate. Development (Cambridge, England) 124 (8):1485–1495 [DOI] [PubMed] [Google Scholar]
- Miller AC, Lyons EL, Herman TG (2009) cis-Inhibition of Notch by endogenous Delta biases the outcome of lateral inhibition. Curr Biol 19 (16):1378–1383 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mohler JD (1956) New mutants report. Drosophila Information Service 30:78–79 [Google Scholar]
- Moloney DJ, Panin VM, Johnston SH, Chen J, Shao L, Wilson R, Wang Y, Stanley P, Irvine KD, Haltiwanger RS, Vogt TF (2000a) Fringe is a glycosyltransferase that modifies Notch. Nature 406 (6794):369–375 [DOI] [PubMed] [Google Scholar]
- Moloney DJ, Shair LH, Lu FM, Xia J, Locke R, Matta KL, Haltiwanger RS (2000b) Mammalian Notch1 is modified with two unusual forms of O-linked glycosylation found on epidermal growth factor-like modules. J Biol Chem 275 (13):9604–9611 [DOI] [PubMed] [Google Scholar]
- Morel V, Lecourtois M, Massiani O, Maier D, Preiss A, Schweisguth F (2001) Transcriptional repression by suppressor of hairless involves the binding of a hairless-dCtBP complex in Drosophila. Curr Biol 11 (10):789–792 [DOI] [PubMed] [Google Scholar]
- Morel V, Schweisguth F (2000) Repression by suppressor of hairless and activation by Notch are required to define a single row of single-minded expressing cells in the Drosophila embryo. Genes Dev 14 (3):377–388 [PMC free article] [PubMed] [Google Scholar]
- Morgan TH, Bridges CB (1916) Sex-linked inheritance in Drosophila. Publs Carnegie Instn 237:66 [Google Scholar]
- Munro S, Freeman M (2000) The notch signalling regulator fringe acts in the Golgi apparatus and requires the glycosyltransferase signature motif DXD. Curr Biol 10 (14):813–820 [DOI] [PubMed] [Google Scholar]
- Nagel AC, Krejci A, Tenin G, Bravo-Patino A, Bray S, Maier D, Preiss A (2005) Hairless-mediated repression of notch target genes requires the combined activity of Groucho and CtBP corepressors. Mol Cell Biol 25 (23):10433–10441 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nam Y, Sliz P, Song L, Aster JC, Blacklow SC (2006) Structural basis for cooperativity in recruitment of MAML coactivators to Notch transcription complexes. Cell 124 (5):973–983 [DOI] [PubMed] [Google Scholar]
- Nasarenko JJ (1930) Ein Fall wahrscheinlicher Verdoppelung eines Chromosomsttlckes bei Drosophila melanogaster. Biol Zentr 50:385–392 [Google Scholar]
- Nichols JT, Miyamoto A, Olsen SL, D’Souza B, Yao C, Weinmaster G (2007) DSL ligand endocytosis physically dissociates Notch1 heterodimers before activating proteolysis can occur. The Journal of cell biology 176 (4):445–458 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nicholson SC, Nicolay BN, Frolov MV, Moberg KH (2011) Notch-dependent expression of the archipelago ubiquitin ligase subunit in the Drosophila eye. Development (Cambridge, England) 138 (2):251–260. 10.1242/dev.054429 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nowell CS, Radtke F (2017) Notch as a tumour suppressor. Nat Rev Cancer 17 (3):145–159. 10.1038/nrc.2016.145 [DOI] [PubMed] [Google Scholar]
- Oberg C, Li J, Pauley A, Wolf E, Gurney M, Lendahl U (2001) The Notch intracellular domain is ubiquitinated and negatively regulated by the mammalian Sel-10 homolog. J Biol Chem 276 (38):35847–35853 [DOI] [PubMed] [Google Scholar]
- Ohlstein B, Spradling A (2006) The adult Drosophila posterior midgut is maintained by pluripotent stem cells. Nature 439 (7075):470–474. 10.1038/nature04333 [DOI] [PubMed] [Google Scholar]
- Okajima T, Irvine KD (2002) Regulation of notch signaling by o-linked fucose. Cell 111 (6):893–904 [DOI] [PubMed] [Google Scholar]
- Okajima T, Reddy B, Matsuda T, Irvine KD (2008) Contributions of chaperone and glycosyltransferase activities of O-fucosyltransferase 1 to Notch signaling. BMC Biol 6:1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Okajima T, Xu A, Lei L, Irvine KD (2005) Chaperone activity of protein O-fucosyltransferase 1 promotes notch receptor folding. Science 307 (5715):1599–1603 [DOI] [PubMed] [Google Scholar]
- Okamura Y, Saga Y (2008) Pofut1 is required for the proper localization of the Notch receptor during mouse development. Mech Dev 125 (8):663–673 [DOI] [PubMed] [Google Scholar]
- Okochi M, Fukumori A, Jiang J, Itoh N, Kimura R, Steiner H, Haass C, Tagami S, Takeda M (2006) Secretion of the Notch-1 Abeta-like peptide during Notch signaling. J Biol Chem 281 (12):7890–7898 [DOI] [PubMed] [Google Scholar]
- Okochi M, Steiner H, Fukumori A, Tanii H, Tomita T, Tanaka T, Iwatsubo T, Kudo T, Takeda M, Haass C (2002) Presenilins mediate a dual intramembranous gamma-secretase cleavage of Notch-1. Embo J 21 (20):5408–5416 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oswald F, Tauber B, Dobner T, Bourteele S, Kostezka U, Adler G, Liptay S, Schmid RM (2001) p300 acts as a transcriptional coactivator for mammalian Notch-1. Mol Cell Biol 21 (22):7761–7774. 10.1128/MCB.21.22.7761-7774.2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oswald F, Winkler M, Cao Y, Astrahantseff K, Bourteele S, Knochel W, Borggrefe T (2005) RBP-Jkappa/SHARP recruits CtIP/CtBP corepressors to silence Notch target genes. Mol Cell Biol 25 (23):10379–10390 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Overgaard MT, Sorensen ES, Stachowiak D, Boldt HB, Kristensen L, Sottrup-Jensen L, Oxvig C (2003) Complex of pregnancy-associated plasma protein-A and the proform of eosinophil major basic protein. Disulfide structure and carbohydrate attachment. J Biol Chem 278 (4):2106–2117 [DOI] [PubMed] [Google Scholar]
- Palermo R, Checquolo S, Giovenco A, Grazioli P, Kumar V, Campese AF, Giorgi A, Napolitano M, Canettieri G, Ferrara G, Schinina ME, Maroder M, Frati L, Gulino A, Vacca A, Screpanti I (2012) Acetylation controls Notch3 stability and function in T-cell leukemia. Oncogene 31 (33):3807–3817. 10.1038/onc.2011.533 [DOI] [PubMed] [Google Scholar]
- Palmer WH, Jia D, Deng WM (2014) Cis-interactions between Notch and its ligands block ligand-independent Notch activity. Elife 3 10.7554/eLife.04415 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pandey A, Li-Kroeger D, Sethi MK, Lee TV, Buettner FF, Bakker H, Jafar-Nejad H (2018) Sensitized genetic backgrounds reveal differential roles for EGF repeat xylosyltransferases in Drosophila Notch signaling. Glycobiology 28 (11):849–859. 10.1093/glycob/cwy080 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Panin VM, Papayannopoulos V, Wilson R, Irvine KD (1997) Fringe modulates Notch-ligand interactions. Nature 387 (6636):908–912 [DOI] [PubMed] [Google Scholar]
- Parks AL, Klueg KM, Stout JR, Muskavitch MA (2000) Ligand endocytosis drives receptor dissociation and activation in the Notch pathway. Development (Cambridge, England) 127 (7):1373–1385 [DOI] [PubMed] [Google Scholar]
- Pei Z, Baker NE (2008) Competition between Delta and the Abruptex domain of Notch. BMC Dev Biol 8:4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perumalsamy LR, Nagala M, Sarin A (2010) Notch-activated signaling cascade interacts with mitochondrial remodeling proteins to regulate cell survival. Proc Natl Acad Sci U S A 107 (15):6882–6887 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Portin P (1975) Allelic negative complementation at the Abruptex locus of Drosophila melanogaster. Genetics 81 (1):121–133 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Posey JE, O’Donnell-Luria AH, Chong JX, Harel T, Jhangiani SN, Coban Akdemir ZH, Buyske S, Pehlivan D, Carvalho CMB, Baxter S, Sobreira N, Liu P, Wu N, Rosenfeld JA, Kumar S, Avramopoulos D, White JJ, Doheny KF, Witmer PD, Boehm C, Sutton VR, Muzny DM, Boerwinkle E, Gunel M, Nickerson DA, Mane S, MacArthur DG, Gibbs RA, Hamosh A, Lifton RP, Matise TC, Rehm HL, Gerstein M, Bamshad MJ, Valle D, Lupski JR, Centers for Mendelian G (2019) Insights into genetics, human biology and disease gleaned from family based genomic studies. Genet Med 21 (4):798–812. 10.1038/s41436-018-0408-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Poulson DF (1937) Chromosomal Deficiencies and the Embryonic Development of Drosophila Melanogaster. Proc Natl Acad Sci U S A 23 (3):133–137 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Presente A, Andres A, Nye JS (2001) Requirement of Notch in adulthood for neurological function and longevity. Neuroreport 12 (15):3321–3325 [DOI] [PubMed] [Google Scholar]
- Presente A, Boyles RS, Serway CN, de Belle JS, Andres AJ (2004) Notch is required for long-term memory in Drosophila. Proc Natl Acad Sci U S A 101 (6):1764–1768. 10.1073/pnas.0308259100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Puente XS, Pinyol M, Quesada V, Conde L, Ordonez GR, Villamor N, Escaramis G, Jares P, Bea S, Gonzalez-Diaz M, Bassaganyas L, Baumann T, Juan M, Lopez-Guerra M, Colomer D, Tubio JM, Lopez C, Navarro A, Tornador C, Aymerich M, Rozman M, Hernandez JM, Puente DA, Freije JM, Velasco G, Gutierrez-Fernandez A, Costa D, Carrio A, Guijarro S, Enjuanes A, Hernandez L, Yague J, Nicolas P, Romeo-Casabona CM, Himmelbauer H, Castillo E, Dohm JC, de Sanjose S, Piris MA, de Alava E, San Miguel J, Royo R, Gelpi JL, Torrents D, Orozco M, Pisano DG, Valencia A, Guigo R, Bayes M, Heath S, Gut M, Klatt P, Marshall J, Raine K, Stebbings LA, Futreal PA, Stratton MR, Campbell PJ, Gut I, Lopez-Guillermo A, Estivill X, Montserrat E, Lopez-Otin C, Campo E (2011) Whole-genome sequencing identifies recurrent mutations in chronic lymphocytic leukaemia. Nature 475 (7354):101–105 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ramain P, Khechumian K, Seugnet L, Arbogast N, Ackermann C, Heitzler P (2001) Novel Notch alleles reveal a Deltex-dependent pathway repressing neural fate. Curr Biol 11 (22):1729–1738 [DOI] [PubMed] [Google Scholar]
- Rampias T, Vgenopoulou P, Avgeris M, Polyzos A, Stravodimos K, Valavanis C, Scorilas A, Klinakis A (2014) A new tumor suppressor role for the Notch pathway in bladder cancer. Nat Med 20 (10):1199–1205. 10.1038/nm.3678 [DOI] [PubMed] [Google Scholar]
- Rana NA, Haltiwanger RS (2011) Fringe benefits: functional and structural impacts of O-glycosylation on the extracellular domain of Notch receptors. Curr Opin Struct Biol 21 (5):583–589 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rand MD, Grimm LM, Artavanis-Tsakonas S, Patriub V, Blacklow SC, Sklar J, Aster JC (2000) Calcium depletion dissociates and activates heterodimeric notch receptors. Mol Cell Biol 20 (5):1825–1835 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ranganathan P, Vasquez-Del Carpio R, Kaplan FM, Wang H, Gupta A, VanWye JD, Capobianco AJ (2011) Hierarchical phosphorylation within the ankyrin repeat domain defines a phosphoregulatory loop that regulates Notch transcriptional activity. J Biol Chem 286 (33):28844–28857. 10.1074/jbc.M111.243600 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rebay I, Fleming RJ, Fehon RG, Cherbas L, Cherbas P, Artavanis-Tsakonas S (1991) Specific EGF repeats of Notch mediate interactions with Delta and Serrate: implications for Notch as a multifunctional receptor. Cell 67 (4):687–699 [DOI] [PubMed] [Google Scholar]
- Rossi D, Trifonov V, Fangazio M, Bruscaggin A, Rasi S, Spina V, Monti S, Vaisitti T, Arruga F, Fama R, Ciardullo C, Greco M, Cresta S, Piranda D, Holmes A, Fabbri G, Messina M, Rinaldi A, Wang J, Agostinelli C, Piccaluga PP, Lucioni M, Tabbo F, Serra R, Franceschetti S, Deambrogi C, Daniele G, Gattei V, Marasca R, Facchetti F, Arcaini L, Inghirami G, Bertoni F, Pileri SA, Deaglio S, Foa R, Dalla-Favera R, Pasqualucci L, Rabadan R, Gaidano G (2012) The coding genome of splenic marginal zone lymphoma: activation of NOTCH2 and other pathways regulating marginal zone development. J Exp Med 209 (9):1537–1551. 10.1084/jem.20120904 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rutten JW, Haan J, Terwindt GM, van Duinen SG, Boon EM, Lesnik Oberstein SA (2014) Interpretation of NOTCH3 mutations in the diagnosis of CADASIL. Expert Rev Mol Diagn 14 (5):593–603. 10.1586/14737159.2014.922880 [DOI] [PubMed] [Google Scholar]
- Sachan N, Mishra AK, Mutsuddi M, Mukherjee A (2013) The Drosophila importin-alpha3 is required for nuclear import of notch in vivo and it displays synergistic effects with notch receptor on cell proliferation. PloS one 8 (7):e68247 10.1371/journal.pone.0068247 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sakamoto K, Chao WS, Katsube K, Yamaguchi A (2005) Distinct roles of EGF repeats for the Notch signaling system. Exp Cell Res 302 (2):281–291. 10.1016/j.yexcr.2004.09.016 [DOI] [PubMed] [Google Scholar]
- Salazar JL, Yamamoto S (2018) Integration of Drosophila and Human Genetics to Understand Notch Signaling Related Diseases. Adv Exp Med Biol 1066:141–185. 10.1007/978-3-319-89512-3_8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sasai Y, Kageyama R, Tagawa Y, Shigemoto R, Nakanishi S (1992) Two mammalian helix-loop-helix factors structurally related to Drosophila hairy and Enhancer of split. Genes Dev 6 (12B):2620–2634. 10.1101/gad.6.12b.2620 [DOI] [PubMed] [Google Scholar]
- Sasaki N, Sasamura T, Ishikawa HO, Kanai M, Ueda R, Saigo K, Matsuno K (2007) Polarized exocytosis and transcytosis of Notch during its apical localization in Drosophila epithelial cells. Genes Cells 12 (1):89–103 [DOI] [PubMed] [Google Scholar]
- Sasamura T, Ishikawa HO, Sasaki N, Higashi S, Kanai M, Nakao S, Ayukawa T, Aigaki T, Noda K, Miyoshi E, Taniguchi N, Matsuno K (2007) The O-fucosyltransferase O-fut1 is an extracellular component that is essential for the constitutive endocytic trafficking of Notch in Drosophila. Development (Cambridge, England) 134 (7):1347–1356 [DOI] [PubMed] [Google Scholar]
- Sasamura T, Sasaki N, Miyashita F, Nakao S, Ishikawa HO, Ito M, Kitagawa M, Harigaya K, Spana E, Bilder D, Perrimon N, Matsuno K (2003) neurotic, a novel maternal neurogenic gene, encodes an O-fucosyltransferase that is essential for Notch-Delta interactions. Development (Cambridge, England) 130 (20):4785–4795 [DOI] [PubMed] [Google Scholar]
- Schnute B, Troost T, Klein T (2018) Endocytic Trafficking of the Notch Receptor. Adv Exp Med Biol 1066:99–122. 10.1007/978-3-319-89512-3_6 [DOI] [PubMed] [Google Scholar]
- Schweisguth F (1999) Dominant-negative mutation in the beta2 and beta6 proteasome subunit genes affect alternative cell fate decisions in the Drosophila sense organ lineage. Proc Natl Acad Sci U S A 96 (20):11382–11386. 10.1073/pnas.96.20.11382 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Selkoe DJ, Hardy J (2016) The amyloid hypothesis of Alzheimer’s disease at 25 years. EMBO Mol Med 8 (6):595–608. 10.15252/emmm.201606210 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shaya O, Sprinzak D (2011) From Notch signaling to fine-grained patterning: Modeling meets experiments. Curr Opin Genet Dev 21 (6):732–739. 10.1016/j.gde.2011.07.007 [DOI] [PubMed] [Google Scholar]
- Shellenbarger DL, Mohler JD (1975) Temperature-sensitive mutations of the notch locus in Drosophila melanogaster. Genetics 81 (1):143–162 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shimizu H, Woodcock SA, Wilkin MB, Trubenova B, Monk NA, Baron M (2014) Compensatory flux changes within an endocytic trafficking network maintain thermal robustness of Notch signaling. Cell 157 (5):1160–1174. 10.1016/j.cell.2014.03.050 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simpson MA, Irving MD, Asilmaz E, Gray MJ, Dafou D, Elmslie FV, Mansour S, Holder SE, Brain CE, Burton BK, Kim KH, Pauli RM, Aftimos S, Stewart H, Kim CA, Holder-Espinasse M, Robertson SP, Drake WM, Trembath RC (2011) Mutations in NOTCH2 cause Hajdu-Cheney syndrome, a disorder of severe and progressive bone loss. Nat Genet 43 (4):303–305. 10.1038/ng.779 [DOI] [PubMed] [Google Scholar]
- Song X, Call GB, Kirilly D, Xie T (2007) Notch signaling controls germline stem cell niche formation in the Drosophila ovary. Development (Cambridge, England) 134 (6):1071–1080. 10.1242/dev.003392 [DOI] [PubMed] [Google Scholar]
- Sportoletti P, Baldoni S, Cavalli L, Del Papa B, Bonifacio E, Ciurnelli R, Bell AS, Di Tommaso A, Rosati E, Crescenzi B, Mecucci C, Screpanti I, Marconi P, Martelli MF, Di Ianni M, Falzetti F (2010) NOTCH1 PEST domain mutation is an adverse prognostic factor in B-CLL. Br J Haematol 151 (4):404–406 [DOI] [PubMed] [Google Scholar]
- Sprinzak D, Lakhanpal A, Lebon L, Santat LA, Fontes ME, Anderson GA, Garcia-Ojalvo J, Elowitz MB (2010) Cis-interactions between Notch and Delta generate mutually exclusive signalling states. Nature 465 (7294):86–90 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stahl M, Uemura K, Ge C, Shi S, Tashima Y, Stanley P (2008) Roles of Pofut1 and O-fucose in mammalian Notch signaling. J Biol Chem 283 (20):13638–13651 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stanley P, Okajima T (2010) Roles of glycosylation in Notch signaling. Curr Top Dev Biol 92:131–164 [DOI] [PubMed] [Google Scholar]
- Stenflo J (1991) Structure-function relationships of epidermal growth factor modules in vitamin K-dependent clotting factors. Blood 78 (7):1637–1651 [PubMed] [Google Scholar]
- Stittrich AB, Lehman A, Bodian DL, Ashworth J, Zong Z, Li H, Lam P, Khromykh A, Iyer RK, Vockley JG, Baveja R, Silva ES, Dixon J, Leon EL, Solomon BD, Glusman G, Niederhuber JE, Roach JC, Patel MS (2014) Mutations in NOTCH1 cause Adams-Oliver syndrome. Am J Hum Genet 95 (3):275–284. 10.1016/j.ajhg.2014.07.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Struhl G, Greenwald I (1999) Presenilin is required for activity and nuclear access of Notch in Drosophila. Nature 398 (6727):522–525. 10.1038/19091 [DOI] [PubMed] [Google Scholar]
- Takeuchi H, Fernandez-Valdivia RC, Caswell DS, Nita-Lazar A, Rana NA, Garner TP, Weldeghiorghis TK, Macnaughtan MA, Jafar-Nejad H, Haltiwanger RS (2011) Rumi functions as both a protein O-glucosyltransferase and a protein O-xylosyltransferase. Proc Natl Acad Sci U S A 108 (40):16600–16605 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takeuchi H, Haltiwanger RS (2010) Role of glycosylation of Notch in development. Semin Cell Dev Biol 21 (6):638–645 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takeuchi H, Schneider M, Williamson DB, Ito A, Takeuchi M, Handford PA, Haltiwanger RS (2018) Two novel protein O-glucosyltransferases that modify sites distinct from POGLUT1 and affect Notch trafficking and signaling. Proc Natl Acad Sci U S A 115 (36):E8395–E8402. 10.1073/pnas.1804005115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tanigaki K, Honjo T (2010) Two opposing roles of RBP-J in Notch signaling. Curr Top Dev Biol 92:231–252 [DOI] [PubMed] [Google Scholar]
- The UniProt C (2017) UniProt: the universal protein knowledgebase. Nucleic Acids Res 45 (D1):D158–D169. 10.1093/nar/gkw1099 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thurmond J, Goodman JL, Strelets VB, Attrill H, Gramates LS, Marygold SJ, Matthews BB, Millburn G, Antonazzo G, Trovisco V, Kaufman TC, Calvi BR, FlyBase C (2019) FlyBase 2.0: the next generation. Nucleic Acids Res 47 (D1):D759–D765. 10.1093/nar/gky1003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tien AC, Rajan A, Bellen HJ (2009) A Notch updated. The Journal of cell biology 184 (5):621–629 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vardar D, North CL, Sanchez-Irizarry C, Aster JC, Blacklow SC (2003) Nuclear magnetic resonance structure of a prototype Lin12-Notch repeat module from human Notch1. Biochemistry 42 (23):7061–7067 [DOI] [PubMed] [Google Scholar]
- Von Halle ES (1965) Localization of E(spl). Drosophila Information Service 40:60 [Google Scholar]
- Vooijs M, Schroeter EH, Pan Y, Blandford M, Kopan R (2004) Ectodomain shedding and intramembrane cleavage of mammalian Notch proteins is not regulated through oligomerization. J Biol Chem 279 (49):50864–50873. 10.1074/jbc.M409430200 [DOI] [PubMed] [Google Scholar]
- Wang J, Al-Ouran R, Hu Y, Kim SY, Wan YW, Wangler MF, Yamamoto S, Chao HT, Comjean A, Mohr SE, Udn, Perrimon N, Liu Z, Bellen HJ (2017) MARRVEL: Integration of Human and Model Organism Genetic Resources to Facilitate Functional Annotation of the Human Genome. Am J Hum Genet 100 (6):843–853. 10.1016/j.ajhg.2017.04.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang J, Liu Z, Bellen HJ, Yamamoto S (2019a) Navigating MARRVEL, a Web-Based Tool that Integrates Human Genomics and Model Organism Genetics Information. J Vis Exp In Press [DOI] [PMC free article] [PubMed]
- Wang J, Mao M, Fazal F, Kim S-Y, Yamamoto S, H.J. B, Liu Z (2019b) Using MARRVEL v1.2 for Bioinformatics Analysis of Human Genes and Variant Pathogenicity. Curr Protoc Bioinformatics In Press [DOI] [PMC free article] [PubMed]
- Wang NJ, Sanborn Z, Arnett KL, Bayston LJ, Liao W, Proby CM, Leigh IM, Collisson EA, Gordon PB, Jakkula L, Pennypacker S, Zou Y, Sharma M, North JP, Vemula SS, Mauro TM, Neuhaus IM, Leboit PE, Hur JS, Park K, Huh N, Kwok PY, Arron ST, Massion PP, Bale AE, Haussler D, Cleaver JE, Gray JW, Spellman PT, South AP, Aster JC, Blacklow SC, Cho RJ (2011) Loss-of-function mutations in Notch receptors in cutaneous and lung squamous cell carcinoma. Proc Natl Acad Sci U S A 108 (43):17761–17766 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y, Spellman MW (1998) Purification and characterization of a GDP-fucose:polypeptide fucosyltransferase from Chinese hamster ovary cells. J Biol Chem 273 (14):8112–8118 [DOI] [PubMed] [Google Scholar]
- Wangler MF, Yamamoto S, Chao HT, Posey JE, Westerfield M, Postlethwait J, Members of the Undiagnosed Diseases N, Hieter P, Boycott KM, Campeau PM, Bellen HJ (2017) Model Organisms Facilitate Rare Disease Diagnosis and Therapeutic Research. Genetics 207 (1):9–27. 10.1534/genetics.117.203067 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wei J, Hemmings GP (2000) The NOTCH4 locus is associated with susceptibility to schizophrenia. Nat Genet 25 (4):376–377. 10.1038/78044 [DOI] [PubMed] [Google Scholar]
- Welshons WJ (1965) Analysis of a gene in drosophila. Science 150 (3700):1122–1129. 10.1126/science.150.3700.1122 [DOI] [PubMed] [Google Scholar]
- Welshons WJ, von Halle ES, Scandlyn BJ (1963) Notch pseudoalleles in Drosophila melanogaster. Int Congr Genet 11 (1):1–2 [Google Scholar]
- Weng AP, Ferrando AA, Lee W, Morris JPt, Silverman LB, Sanchez-Irizarry C, Blacklow SC, Look AT, Aster JC (2004) Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia. Science 306 (5694):269–271 [DOI] [PubMed] [Google Scholar]
- Wesley CS (1999) Notch and wingless regulate expression of cuticle patterning genes. Mol Cell Biol 19 (8):5743–5758 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wharton KA, Johansen KM, Xu T, Artavanis-Tsakonas S (1985a) Nucleotide sequence from the neurogenic locus notch implies a gene product that shares homology with proteins containing EGF-like repeats. Cell 43 (3 Pt 2):567–581 [DOI] [PubMed] [Google Scholar]
- Wharton KA, Yedvobnick B, Finnerty VG, Artavanis-Tsakonas S (1985b) opa: a novel family of transcribed repeats shared by the Notch locus and other developmentally regulated loci in D. melanogaster. Cell 40 (1):55–62 [DOI] [PubMed] [Google Scholar]
- Wilson JJ, Kovall RA (2006) Crystal structure of the CSL-Notch-Mastermind ternary complex bound to DNA. Cell 124 (5):985–996 [DOI] [PubMed] [Google Scholar]
- Witt CM, Won WJ, Hurez V, Klug CA (2003) Notch2 haploinsufficiency results in diminished B1 B cells and a severe reduction in marginal zone B cells. J Immunol 171 (6):2783–2788. 10.4049/jimmunol.171.6.2783 [DOI] [PubMed] [Google Scholar]
- Wu G, Lyapina S, Das I, Li J, Gurney M, Pauley A, Chui I, Deshaies RJ, Kitajewski J (2001) SEL-10 is an inhibitor of notch signaling that targets notch for ubiquitin-mediated protein degradation. Mol Cell Biol 21 (21):7403–7415 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu A, Haines N, Dlugosz M, Rana NA, Takeuchi H, Haltiwanger RS, Irvine KD (2007) In vitro reconstitution of the modulation of Drosophila Notch-ligand binding by Fringe. J Biol Chem 282 (48):35153–35162 [DOI] [PubMed] [Google Scholar]
- Xu A, Lei L, Irvine KD (2005) Regions of Drosophila Notch that contribute to ligand binding and the modulatory influence of Fringe. J Biol Chem 280 (34):30158–30165 [DOI] [PubMed] [Google Scholar]
- Xu T, Artavanis-Tsakonas S (1990) deltex, a locus interacting with the neurogenic genes, Notch, Delta and mastermind in Drosophila melanogaster. Genetics 126 (3):665–677 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu T, Caron LA, Fehon RG, Artavanis-Tsakonas S (1992) The involvement of the Notch locus in Drosophila oogenesis. Development (Cambridge, England) 115 (4):913–922 [DOI] [PubMed] [Google Scholar]
- Xu T, Rebay I, Fleming RJ, Scottgale TN, Artavanis-Tsakonas S (1990) The Notch locus and the genetic circuitry involved in early Drosophila neurogenesis. Genes Dev 4 (3):464–475. 10.1101/gad.4.3.464 [DOI] [PubMed] [Google Scholar]
- Xu X, Choi SH, Hu T, Tiyanont K, Habets R, Groot AJ, Vooijs M, Aster JC, Chopra R, Fryer C, Blacklow SC (2015) Insights into Autoregulation of Notch3 from Structural and Functional Studies of Its Negative Regulatory Region. Structure 23 (7):1227–1235. 10.1016/j.str.2015.05.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamamoto S, Charng WL, Bellen HJ (2010) Endocytosis and intracellular trafficking of Notch and its ligands. Curr Top Dev Biol 92:165–200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamamoto S, Charng WL, Rana NA, Kakuda S, Jaiswal M, Bayat V, Xiong B, Zhang K, Sandoval H, David G, Wang H, Haltiwanger RS, Bellen HJ (2012) A mutation in EGF repeat-8 of Notch discriminates between Serrate/Jagged and Delta family ligands. Science 338 (6111):1229–1232. 10.1126/science.1228745 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamamoto S, Jaiswal M, Charng WL, Gambin T, Karaca E, Mirzaa G, Wiszniewski W, Sandoval H, Haelterman NA, Xiong B, Zhang K, Bayat V, David G, Li T, Chen K, Gala U, Harel T, Pehlivan D, Penney S, Vissers L, de Ligt J, Jhangiani SN, Xie Y, Tsang SH, Parman Y, Sivaci M, Battaloglu E, Muzny D, Wan YW, Liu Z, Lin-Moore AT, Clark RD, Curry CJ, Link N, Schulze KL, Boerwinkle E, Dobyns WB, Allikmets R, Gibbs RA, Chen R, Lupski JR, Wangler MF, Bellen HJ (2014a) A drosophila genetic resource of mutants to study mechanisms underlying human genetic diseases. Cell 159 (1):200–214. 10.1016/j.cell.2014.09.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamamoto S, Schulze KL, Bellen HJ (2014b) Introduction to Notch signaling. Methods Mol Biol 1187:1–14. 10.1007/978-1-4939-1139-4_1 [DOI] [PubMed] [Google Scholar]
- Yang G, Zhou R, Zhou Q, Guo X, Yan C, Ke M, Lei J, Shi Y (2019) Structural basis of Notch recognition by human gamma-secretase. Nature 565 (7738):192–197. 10.1038/s41586-018-0813-8 [DOI] [PubMed] [Google Scholar]
- Yao D, Huang Y, Huang X, Wang W, Yan Q, Wei L, Xin W, Gerson S, Stanley P, Lowe JB, Zhou L (2011) Protein O-fucosyltransferase 1 (Pofut1) regulates lymphoid and myeloid homeostasis through modulation of Notch receptor ligand interactions. Blood 117 (21):5652–5662 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ye Y, Lukinova N, Fortini ME (1999) Neurogenic phenotypes and altered Notch processing in Drosophila Presenilin mutants. Nature 398 (6727):525–529. 10.1038/19096 [DOI] [PubMed] [Google Scholar]
- Zhang N, Martin GV, Kelley MW, Gridley T (2000) A mutation in the Lunatic fringe gene suppresses the effects of a Jagged2 mutation on inner hair cell development in the cochlea. Curr Biol 10 (11):659–662 [DOI] [PubMed] [Google Scholar]
- Zhou X, Li H, Guo S, Wang J, Shi C, Espitia M, Guo X, Wang Q, Liu M, Assassi S, Reveille JD, Mayes MD (2019) Associations of Multiple NOTCH4 Exonic Variants with Systemic Sclerosis. J Rheumatol 46 (2):184–189. 10.3899/jrheum.180094 [DOI] [PubMed] [Google Scholar]