Next Article in Journal
The Lewis Pair Polymerization of Lactones Using Metal Halides and N-Heterocyclic Olefins: Theoretical Insights
Next Article in Special Issue
HIV Tat/P-TEFb Interaction: A Potential Target for Novel Anti-HIV Therapies
Previous Article in Journal
Inhibition of Cytochrome P450 Activities by Extracts of Hyptis verticillata Jacq.: Assessment for Potential HERB-Drug Interactions
Previous Article in Special Issue
Thymoquinone Inhibits the Migration and Invasive Characteristics of Cervical Cancer Cells SiHa and CaSki In Vitro by Targeting Epithelial to Mesenchymal Transition Associated Transcription Factors Twist1 and Zeb1
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Taming the Notch Transcriptional Regulator for Cancer Therapy

1
Laboratory of Cancer Cell Biology, Candiolo Cancer Institute-FPO, IRCCS, Str. Prov. 142, 10060 Candiolo, TO, Italy
2
Department of Medical, Surgical and Health Sciences, University of Trieste, 34127 Trieste, Italy
3
Department of Oncology, University of Torino, c/o IRCCS, S.P. 142, 10060 Candiolo, TO, Italy
4
Cardiovascular Biology Laboratory, International Centre for Genetic Engineering and Biotechnology (ICGEB), 34149 Trieste, Italy
*
Authors to whom correspondence should be addressed.
Molecules 2018, 23(2), 431; https://doi.org/10.3390/molecules23020431
Submission received: 22 January 2018 / Revised: 12 February 2018 / Accepted: 13 February 2018 / Published: 15 February 2018
(This article belongs to the Special Issue Transcription Factors as Therapeutic Targets)

Abstract

:
Notch signaling is a highly conserved pathway in all metazoans, which is deeply involved in the regulation of cell fate and differentiation, proliferation and migration during development. Research in the last decades has shown that the various components of the Notch signaling cascade are either upregulated or activated in human cancers. Therefore, its downregulation stands as a promising and powerful strategy for cancer therapy. Here, we discuss the recent advances in the development of small molecule inhibitors, blocking antibodies and oligonucleotides that hinder Notch activity, and their outcome in clinical trials. Although Notch was initially identified as an oncogene, later studies showed that it can also act as a tumor suppressor in certain contexts. Further complexity is added by the existence of numerous Notch family members, which exert different activities and can be differentially targeted by inhibitors, potentially accounting for contradictory data on their therapeutic efficacy. Notably, recent evidence supports the rationale for combinatorial treatments including Notch inhibitors, which appear to be more effective than single agents in fighting cancer.

1. Diversity of Notch Signaling Mechanisms

This is the centennial of the discovery of the Notch gene, which was first identified by Thomas H. Morgan and colleagues in 1917 in fruit flies, where spontaneous mutations in a specific locus of the X chromosome produced notches at the wing margin [1,2]. Molecular genetic studies followed, but it took until the 1980s to accomplish the actual cloning of Drosophila Notch gene by Artavanis-Tsakonas, Young, and colleagues [3,4]. Notably, the Notch pathway is highly conserved from sea urchins to humans. In contrast to flies that have only one gene, there are four Notch receptors in mammals, Notch1 to Notch4. Notch signaling regulates differentiation, proliferation, apoptosis, migration, and angiogenesis, as well as stem cell growth and survival during development and disease [5,6,7,8,9].
Notch proteins are single pass transmembrane heterodimeric receptors. They are synthesized as single-chain precursors that undergo processing and modification in the endoplasmic reticulum and the Golgi apparatus to produce the mature forms [10]. In particular, Notch precursor undergoes a first proteolytic cleavage (S1) in the Golgi to form a heterodimer, which represents the mature form of the receptor. This heterodimer sits on the plasma membrane and is composed of a large extracellular domain (N-ECD) and a membrane-tethered intracellular domain (N-ICD). Notch receptors are activated by ligands of the Delta-like (Dll) and Jagged (Jag) families that are exposed on the surface of adjacent cells, thereby binding the N-ECD in trans and acting in a juxtracrine manner. In total, there are five Notch ligands that belong to Delta like ligand (DLL) and Jagged families—Dll1, Dll3, Dll4, Jag1, and Jag2. Dll and Jagged ligands contain Delta-Serrate-Lag2 (DSL) domain and EGF (Epidermal Growth Factor) repeats that interact with EGF-like repeats found in the Notch ectodomain to activate the signaling cascade. Interactions between Notch and its ligands in cis, instead, are reported to be inhibitory [11]. The glycosylation of EGF repeats is an important modulator of Dll vs Jagged signaling; for instance, Fringe glycosyltransferases (that add GlcNAc to O-fucose on Notch EGF repeats) can enhance the binding of Dll ligands, compared to Jagged, to the Notch receptor [12]. Upon ligand binding, Notch is sequentially cleaved at the plasma membrane by two other proteases. Extracellular S2 cleavage is due to the tumor necrosis factor-alpha converting enzyme (TACE), whereas S3 intramembrane cleavage is operated by a gamma secretase, which catalyzes the release of the intracellular domain of Notch (N-ICD) into the cytoplasm. N-ICD then translocates inside the nucleus and forms a complex with RBPjk/CSL (CBF1/Su(H)/Lag-1) [13]. RBPjk is a DNA-binding transcription factor, which acts in a large complex with other proteins. It is the prime effector of Notch-associated functions. Normally RBPjk acts as a repressor of Notch transcriptional target genes due to its association with SMRT corepressor, CIR, and histone deacetylase-1 (HDAC1). N-ICD binding disrupts this inhibitory complex, releasing the repressor proteins and fostering the formation of a new complex of RBPjk with SKIP, MAML1, PCAF, and GCN5, which is able to induce gene expression. Notch signaling regulates a diverse set of genes that impinge on several cellular functions [6]. Among the known Notch-RBPjk targets are the genes belonging to the HES and HEY families, NF-κB1, NF-κB2, STAT6, p27/Kip1, ErbB2, CyclinD1, cMyc, p21.
DSL-ligand- and RBPjk- independent Notch-initiated signaling has also been reported, referred to as “non-canonical” Notch pathway [14,15]. While the canonical pathway plays a major and well defined role in stem cell maintenance and cell fate determination during development, and it regulates several cancer hallmarks in tumorigenesis, the functions of the non-canonical pathway are poorly characterized [14]. Neither the targets nor the mediators of the non-canonical pathway are clearly defined; however, N-ICD has been reported to associate with other transcription factors distinct from RBPjk, such as SMAD3, YY1, HIF1α, and NF-κB [16,17,18,19]. In one study, it was shown that IL-6, an important pro-inflammatory cytokine in tumors, was upregulated by Notch1-ICD in breast cancer cells with p53 mutation or deficiency, independent of RBPjk [20]; IKKα and IKKβ were implicated in this mechanism, but whether they can form a complex with N-ICD needs to be elucidated. Several studies illustrated the crosstalk between hypoxia-induced HIF and Notch transcription factors; for instance, HIF1α was found to interact and stabilize the N-ICD, leading to enhancement of the canonical Notch pathway [21,22,23]. On the other hand, it was shown that the negative regulator FIH-1 (Factor Inhibiting HIF1) can bind and functionally inactivate N-ICD by hydroxylation [18]. In addition, RBPjk–independent Notch signaling was reported to activate the PI3K pathway in cervical cancer cell lines, implicating the activity of Deltex1, an E3-ubiquitin ligase [24].

2. Diversity of Notch Signaling Activities in Cancer

Ellisen and coworkers were the first to associate Notch with cancer; they reported that in T-cell acute lymphocytic leukemia (T-ALL), Notch1 gets constitutively activated due to chromosomal translocation t(7;9)(q34; q34.3), found in about 1% of T-ALL patients [21]. Later studies showed that more than 50% of T-ALLs actually show activating Notch1 mutations in the heterodimerization (HD) and ‘proline, glutamic acid, serine, threonine-rich’ (PEST) domains [22]. Remarkably, the majority of solid human cancers bear deregulated expression of Notch pathway genes, rather than genetic changes [23]. Notch1 is the best-studied member of the family, and it is often overexpressed in human breast, colorectal, lung, pancreas, and prostate cancers [6]. Notch1 expression is induced by hypoxia, which is a common event in growing tumors [24,25]. Moreover, Notch signaling is activated downstream of a range of pathways deregulated in cancer, such as cMyc, p53, PI3K, or RAS [26,27,28,29]. Several miRNAs can furthermore control the Notch pathway, e.g., miR34a and miR326 have been shown to target both Notch1 and Notch2 in gliomas and pancreatic cancer [30,31,32], and Notch3 3′ UTR can be targeted by miR206 [33].
Notch activity regulates several cancer hallmarks, such as cancer cell survival, proliferation, migration, invasion, and metastasis, as it transcriptionally modulates a range of signaling pathways. For instance, the ERK pathway and cyclinD1 seem to mediate Notch1-dependent control of cell survival and proliferation [34,35,36]. Moreover, Notch regulates mTOR/Akt and NF-κB pathways, eventually inhibiting apoptosis in breast cancer cells [37,38,39]. Notch1 and its ligands were found to be overexpressed in prostate cancer compared to normal tissue [40]. Notably, prostate cancers are commonly characterized by inactivation of the tumor suppressor PTEN [41,42], which is negatively regulated by Hes1, a well-known effector of canonical Notch signaling [43]. In addition, Kwon and coworkers showed that in PTEN null mice Notch signaling was required for metastatic dissemination, through the induction of epithelial to mesenchymal transition (EMT) in a FOXC2-dependent manner [40]. Indeed, loss of E-cadherin and EMT are key events in prostate cancer metastasis [44]. EMT is characterized by increased expression of a series of transcription factors (including Snail, Slug, Twist, Zeb1, and Zeb2), and the Notch pathway induces their expression in prostate as well as in breast, pancreatic, and colon cancers [45,46,47]. Notably, in prostate cancer cells, Notch was also shown to regulate an EMT-like phenotype through the upregulation of the semaphorin receptor PlexinD1, impinging on SLUG [48]. Moreover, PlexinD1 expression correlates with Notch1 and Notch3 in different cancer types [48], and its role in metastatic tumor progression has been further demonstrated in colon cancer, melanoma, and ovarian cancer [49,50]. In apparent contradiction with these data, in endothelial cells Notch appears to negatively regulate PlexinD1 expression [51]. Thus, the Notch-PlexinD1 axis seems to exert a range of effects in diverse cells, warranting further studies to fully define its role in cancer.
Notably, in certain contexts, Notch signaling has been shown to be tumor suppressive, such as in head and neck carcinomas, and in pancreatic cancer [52,53]. In squamous carcinomas of skin and lung, about 75% of patients bear loss of function mutations in Notch1 or Notch2 [54]. In fact, Notch upregulates p21/Cip1/WAF expression in keratinocytes, which supports its tissue-specific tumor-suppressive function in the skin [55]. A similar pattern of Notch1/2/3 functional inactivation has been reported in about 40% of bladder cancers [56]. In pancreatic adenocarcinoma, acute myeloid leukemia, and angiosarcomas, the role of Notch1 remains controversial [52,57].
The functional role of the other family members, Notch2, Notch3, and Notch4 is less understood. For instance, Ortica et al. had comparatively tested the function of the four Notch receptors in mouse embryonic cells by over-expressing constitutively active forms, and verified their diversity in controlling cell proliferation and differentiation fate [58]. Notably, in colorectal cancer Notch2 levels are decreased and similar results were obtained in thyroid and ovarian cancers, suggesting a tumor suppressor function for this family member [59,60]. On the other hand, Notch3 is constitutively activated in one third of basal-like breast cancers, and a Notch3 specific antagonistic antibody, anti-N3.A4, hampered the growth of orthotopic HCC1143 breast cancer xenografts in mice [61]. In prostate cancer, Notch3 overexpression was associated with higher Gleason score and a high proliferative gene expression signature [62]. In hepatocellular cancer, both Notch1 and Notch3 levels correlated with tumor grade, invasion, and metastasis [63]. In oral squamous cell carcinomas, Notch3 signaling is activated in stromal fibroblasts, in turn eliciting tumor angiogenesis [64]. Besides NOTCH1, also NOTCH3 gene was found to be activated by mutations in T-ALL [65]; moreover, expression analysis of T-ALL cells revealed a common oncogenomic program triggered by both activated Notch oncogenes via RBPjk [65,66]. In other studies, Notch3 was instead linked to tumor suppressive activity, e.g., its overexpression in breast and melanoma cell lines was found to increase p21 and thereby inhibit cell proliferation and induce cell senescence [67]. Zhang et al. recently reported that Notch3 induces the tumor suppressor WWC1/Kibra, a regulator of the Hippo pathway, thus inhibiting EMT in breast cancer cells [68]. Recent data seem to support a pro-tumorigenic function of Notch4. For example, in triple negative breast cancer (TNBC) cell line MDA-MB-231, Notch4 overexpression induces proliferation and invasion and, conversely, its downregulation inhibits proliferation [69]. In another model of breast cancer—MCF7 cells—increased expression of Notch4 elicited EMT and invasiveness [70]. Notch4 is also upregulated in pancreatic cancer cell lines compared to non-transformed cells and its inhibition impairs viability, migration, and invasion [71]. Notch4 activity was also associated with EMT gene expression signature in melanoma cells and held responsible for increased metastasis [72].
An additional layer of complexity in the role of Notch signaling in cancer is added by the recent findings that the expression of several Notch target genes remains upregulated even after RBPjk depletion due to epigenetic changes (enrichment of H3K4me3 and H4ac marking the active promoters) [73]. Notably, there is controversy about the role of RBPjk in cancer: in fact, consistent with its transcriptional repressor function, RBPjk depletion can promote tumorigenesis [73,74]; however, RBPjk was also found highly expressed in glioblastomas, and its targeting decreased self-renewal of brain tumor-initiating cell and tumor formation [75].

3. Advances in Targeting of Notch Signaling by Small Molecule Inhibitors

The development of small molecules able to target signaling molecules, specifically active in tumor cells, has exponentially increased over the last decades. It is relatively easier to develop targeted drugs against catalytic sites, such as for oncogenic protein kinases. Instead, the identification of appropriate targeting strategies for non-enzymatic molecules is more challenging. Based on the structure of Notch receptors and their mechanism of activation, several types of inhibitors have been generated, which are able to inhibit either the binding of the ligands or the gamma secretase-dependent proteolytic cleavage of the receptor [76] (Figure 1). These inhibitors have been tested either alone or in combination, and some of them entered clinical trials with alternating success. Gamma secretase inhibitors (GSI), which were initially developed to prevent the formation of amyloid deposits in Alzheimer’s disease, have been repurposed to block other targets of the same enzymes, including the Notch family receptors that, as mentioned, are often highly expressed in cancer. Gamma secretase is in fact a complex of proteins, including APH1, PEN2, Nicastrin, and presenilin. It has more than 100 type-I membrane protein targets, including Notch ligands, ErbB4, Syndecan, and CD44 [77,78,79]. The gamma secretase complex cleaves within the intramembrane region of its protein substrates and is very promiscuous regarding the target sequence. More than 100 GSI have been synthesized so far, with different affinity, specificity, and IC50 values, falling into 3 categories: (i) peptide isosteres, (ii) azepines, and (iii) sulfonamides [80]. GSI are either transition state analogs of the aspartyl proteinase active site (namely, they bind competitively to the catalytic site of presenilins) or non-transition state inhibitors, which bind in a site other than the active site (i.e., at the dimerization interface of the gamma secretase complex). An early generation non-transition state analogue is DAPT (N-[N-(3,5-Difluorophenacetyl-l-alanyl)]-S-phenylglycine t-Butyl Ester) [81]. Later, other compounds were generated, with 100-fold stronger activity, such as LY411575, LY450139, RO4929097, and BMS906024 (Figure 2). RO4929097 has been shown to extend survival in an intracranial mouse glioma model [82], whereas BMS906024 is the only GSI able to efficiently inhibit all four Notch receptors, as well as APP [83]. The non-transition state analogs, like DAPT, are supposed to be more effective than the transition state inhibitors, as substrate docking into the enzyme interior can hinder inhibitor binding in the transition state. The overall efficacy of GSIs in cancer is still under debate, as several studies have been performed in diverse models, which makes it difficult to draw univocal conclusions (Figure 3). MCF7 breast cancer cells stably expressing N1-ICD underwent EMT and grew faster in a xenograft model and, conversely, treatment of SKBR3 xenografts with DAPT modestly reduced tumor growth [84,85]. The blockade of Notch pathway by GSI-18 in mouse GBM xenografts depleted CD133-positive stem-like cells, reduced tumor growth, and significantly promoted survival [86]. These and other studies paved the way for clinical trials, in which many GSIs have been administered orally (Figure 4). The current clinical trials are focused on RO4929097, MK0752, and PF03084014. In particular, RO4929097 has been tested in phase I and II trials in patients with advanced solid tumors. In patients with colorectal carcinoma it did not show any positive effects, but instead mild toxicity with nausea and vomiting [87]. An additional, frequent adverse effect of GSI is diarrhea, associated with excess secretory goblet cells in the intestine, possibly because Notch inhibition causes an imbalance in cell fate differentiation [88]. Due to the lack of clinical benefit in multiple studies, its clinical use was terminated prematurely [89]. MK0752 has also been tested in phase I and II clinical trials, showing partial clinical benefit in a cohort of 103 patients with advanced solid tumors; one patient with anaplastic astrocytoma showed complete response lasting for more than one year [90]. Another shortcoming of GSI is their lack of effectiveness upon the onset of drug-resistance mechanisms. This has been primarily reported in T-ALL, which showed lack of benefit from GSI despite the constitutive Notch activity. GSI-resistant T-ALL cells were found to harbor mutational loss of PTEN, and thereby aberrant activation of the Akt pathway, increased glycolysis, and carbon metabolism [43]. T-ALL cell lines were also found to harbor FBW7 mutations leading to residual Notch signaling, which contributed to GSI resistance [91]. Furthermore, GSI-resistant T-ALL cells can maintain upregulated Myc expression independently of Notch, due to Brd4 activity [92].
As mentioned before, a major limitation of molecules targeting multiple Notch receptors stems from the different functions exerted by the different Notch members. They can all act as either oncogenes or tumor suppressor genes, depending on the context. Indeed, in some instances, GSI therapy was found to aggravate skin cancer [93,94]. It has also been reported that chronic treatment with Notch inhibitors leads to the formation of vascular tumors [95], consistent with the fact that Notch regulates angiogenesis, and its inhibition leads to exuberant vessel sprouting and defective maturation. Notably, only a few studies have investigated the long-term consequences of Notch inhibition. In transgenic mice undergoing progressive loss of Notch1 with age, a shorter life-span of about 10 months was observed due to widespread vascular tumors [96], such as liver hemangiomas, due to excessive proliferation of endothelial cells. In mice, GSI therapy was associated with immunosuppression and inhibition of stem cell renewal in normal tissues [97]. Intermittent regimens of systemic Notch inhibition, i.e., multiple drug cycles separated by a drug holiday, are currently being tested in an attempt to increase efficacy and minimize toxicity of GSIs. Specific targeting of the Notch pathway is also potentially achievable by tackling the N-ICD/RBPjk/MAML ternary complex (NTC) at nuclear level. For example, by using an innovative combination of molecular docking of small molecules with NTC and proximity-based AlphaScreen technology, Astudillo et al. designed a small molecule inhibitor of NTC, called Inhibitor of Mastermind recruitment-1 (IMR-1) [98]. IMR-1 selectively inhibited Notch-dependent transcriptional activation by interfering with Maml1 recruitment to the N-ICD/RBPjk complex, and impaired tumor growth in patient-derived xenograft (PDX) models.

4. Combinatorial Treatments with Notch Inhibitors

Due to the lack of success of GSIs as single agents, there is now growing interest on their use in association with standard cancer treatments, including hormonal, radiation, chemotherapy, or targeted inhibitors (see Figure 1). Interestingly, cancer treatment with a range of chemotherapeutic agents, such as paclitaxel, docetaxel, cisplatin, oxaliplatin, doxorubicin, and gemcitabine independently resulted in the upregulation of the Notch pathway [99]. This clearly supports the rationale for combination therapies, since the Notch pathway may represent a survival mechanism activated to endure drug treatment. For instance, in head and neck squamous cell carcinoma (HNSCC), cisplatin-resistant tumors showed high expression of Notch1 and a good response to GSIs [100]. Similarly, in colorectal and ovarian cancers, the combination of cisplatin and GSIs was more effective than either monotherapy. In colon cancer, 5-FU and oxaliplatin induced Notch activity, which was curbed by GSI association [101]. In pancreatic cancer, in which gemcitabine chemotherapy is a standard treatment, drug-resistant tumors upregulated Notch2, Notch3, and Jag1, while the inhibition of Notch3 and gemcitabine induced apoptosis [102]. Notch pathway inhibitors have been shown to synergize with DNA damaging agents like doxorubicin in a variety of breast cancer cell lines [103,104]. TNBC is a good example of the use of combination therapy. TNBC and basal-like breast cancers often express particularly high levels of both EGFR and Notch. Dong and coworkers showed that the combined blockage of both pathways by Gefitinib and Compound-E is highly effective in suppressing HCC1806 tumor growth in a xenograft mouse model [105]. Moreover, Pandya et al. showed that a combination therapy with anti-Her2 and GSI prevents drug-resistance and tumor recurrence of HER2-overexpressing xenografts [106]. In prostate cancer, in which anti-androgen or chemotherapy is often the standard treatment, combination of docetaxel and the Notch inhibitor PF03084014 was effective in preclinical models, resulting in reduced cancer cell EMT, prosta-spheres formation, and tumor microvessel density in vivo [107]. Jin and coworkers showed a synergistic effect of the Notch inhibitor MRK003 with the Akt inhibitor MK2206, mainly impacting on invasion rather than on cancer cell proliferation [108]. In addition, in KRAS-driven lung adenocarcinomas, the co-inhibition of DDR1 (by dasatinib) and Notch pathway (by demcizumab) curbed tumor cell proliferation with additive therapeutic benefit [109]. Synergy of GSIs with the inhibition of Bcl-2 and Bcl-xL by ABT737 was also reported in multiple myeloma, in which the combined therapy resulted in the increased activity of Bak, Bax, and release of cytochrome-c, with no effects on peripheral blood mononuclear cells [110]. In T-ALL, in which Notch itself is activated in about 50% of cases, there is a population of cells that is GSI-tolerant and expands even in the presence of Notch1 inhibitors. Interestingly, increased expression of Myc in more than 50% T-ALL is reportedly due to Notch activation [26,111]. In fact, the association of GSI with a BET-bromodomain inhibitor (JQ1) downregulating Myc and Myc target genes showed increased efficacy in inducing apoptosis in T-ALL [107].

5. Inhibition of Notch Signaling by Blocking Antibodies and Decoys

As discussed above, GSIs inhibit not only Notch receptor activation but also other gamma secretase targets, which often lead to adverse effects and ambiguous conclusions regarding safety and efficacy. The development of therapeutic antibodies, capable of specifically inhibiting Notch family members, stands as a promising strategy to overcome this limitation. Blocking antibodies to Notch can be divided into two groups: (a) those directed to the ‘negative regulatory region’ (NRR) that enable ADAM mediated cleavage, e.g., raised against Notch1–3 by phage display technology [112]; and (b) antibodies that block receptor-ligand interaction by hindering EGF repeats. Humanized antibodies against Notch1 or Notch2/3 have been generated and entered phase I and II trials. Moreover, a cross-reactive human monoclonal Notch2 and Notch3 antagonist, OMP-59R5 (Tarextumab), is effective in reducing cancer cell proliferation and the growth of breast, lung, ovarian, and pancreatic cancers [113]. OMP-59R5 was also found to be well tolerated in advanced pancreatic cancer patients and showed efficacy in combination with gemcitabine and Nab-paclitaxel [113,114]. Notably, the use of antibodies selectively targeting Notch1 and sparing Notch2 has been shown to prevent major adverse and toxic effects related with pan-Notch inhibition [115]. Notch ligands can also be targeted by blocking molecules. For instance, anti-Dll4 antibodies have been tested in pre-clinical and clinical trials. The humanized anti-Dll4 mAb OMP21M18 caused inhibition of tumor growth in PDX models and is currently being tested in several clinical trials, either as a single agent or in combination with other drugs [116]. A neutralizing anti-DLL4 humanized phage antibody, YW152F, was shown to both specifically block Notch-DLL4 signaling and hamper the growth of MDA-MB-435 cells, implanted in the mouse mammary fat pad [117]. Combination of DLL4-blockade by either Dll4-Fc or anti-DLL4 antibodies (REGN1035, REGN421) with VEGF-trap Aflibercept showed stronger efficacy in reducing tumor burden compared to monotherapy, by blocking angiogenesis and eliciting cancer cell apoptosis in murine tumors [118]. Soluble decoys blocking Notch signaling have also been tested. For instance, Funahashi et al. generated a soluble form of the Notch1 ectodomain, which competes with Dll1 and Jag1 binding to the receptor. This molecule was particularly effective in inhibiting VEGF-induced angiogenesis in normal skin and neo-angiogenesis in tumor models [119]. Kuramoto and coworkers generated a soluble form of Dll4, called Dll4-Fc, which suppressed liver metastasis of small cell lung cancer (SCLC) cells expressing high levels of Dll4 [120]. Intriguingly, the non-canonical Notch ligand DLK1 could also compete with canonical ligands of the DSL type to block the canonical signaling pathway [121]. The Notch receptor is a membrane-tethered protein, and physiologically no soluble forms have been reported to compete with its activity. The extracellular EGF-like11-13 domains in the Notch receptor have been artificially fused to the IgG1 Fc domain to form a soluble protein capable to trap Notch ligands. Using this approach, Klose et al. showed that Notch1-EGF11-13-Fc, which showed preferential binding to Jag1, inhibited in vitro tube formation and tip cell formation in the retinal angiogenesis assay [122]. Recently, Kangsamaskin et al. generated Notch decoys specific for the unique region of ligand-receptor interaction [123]. They showed that the N110-24 decoy effectively inhibited angiogenic sprouting in the mouse retina, as well as tumor growth, by specifically blocking Jag1/Jag2 mediated Notch signaling. In contrast, the N11-13 decoy led to vessel hypersprouting in the same models by interfering with Dll1/Dll4 mediated Notch signaling. An additional approach, less explored so far, is to prevent the assembly of Notch transcription complex, composed of N-ICD, RBPjk, and MAML. For instance, Mollering et al. generated 16-amino acid long peptides, based on MAML sequence motif required for interaction with N-ICD and RBPjk, called ’stapled α-helical peptides derived from MAML1′ (SAHM). Indeed, SAHM1 prevented the assembly of active transcription complex by binding to N-ICD and RBPjk, thereby inhibiting the recruitment of the MAML1 co-activator [124]. SAHM1 peptide treatment suppressed Notch target genes in the T-ALL cell line KOPT-K1 and concomitantly halted proliferation and leukemia progression. Clearly, molecules interfering with Notch-activated transcriptional complex may be expected to recapitulate pan-Notch signaling inhibition, including some of the reported adverse effects. Moreover, the therapeutic use of peptides in human patients poses pharmacokinetic issues to be addressed.

6. Recent Trends in Notch Targeting by Oligonucleotide-Based Methods

Antisense oligonucleotides are short stretches of chemically synthesized DNA or RNA able to target genes of interests [125]. Due to high sequence specificity, they are expected to show less off-target effects compared to other therapeutic approaches. As nucleic acids do not easily enter inside the cells, they are generally chemically modified to increase their capacity to penetrate cell membranes. Nakazawa et al. reported that Notch signaling regulates the abnormal proliferation and differentiation of synoviocytes in Rheumatoid Arthritis. The proliferation of synoviocytes could be blocked by both Notch1 antisense oligonucleotides and the gamma secretase inhibitor MW167 [126]. Moreover, in the SCLC cell line NCI-H82, the use of Notch1-phosphorothioate antisense oligonucleotides reduced the expression of Notch target Hes1 [127]. To modulate angiogenesis, Zimrin and coworkers generated an antisense Jagged oligomer, which potentiated FGF2-induced collagen invasion by endothelial cells [128]. Very few studies have exploited these tools in vivo in preclinical models, with one limitation being the poor efficiency of intracellular delivery upon systemic administration. Notably, the systemic expression of Notch1 antisense oligonucleotides (NAS) under mouse mammary tumor virus long terminal repeat promoter was found to significantly reduce Notch activity in tissues, but also showed several Notch-associated defects on learning and memory [129,130]. Thus, an accurate titration of the systemic delivery of NAS is warranted, with regard to their application for cancer therapy.

7. Conclusions and Future Perspectives

The Notch transcriptional regulator lies at the intersection of multiple signaling pathways, which control cell stemness, differentiation, survival, proliferation, migration, and invasion during development and disease. In most instances, the inhibition of Notch signaling has been shown to block cancer cell growth and angiogenesis. In fact, the deregulated expression of Notch pathway components in human cancers supports the therapeutic use of GSIs. Recent studies tend to indicate that these drugs may be more effective in combination with chemotherapy or other targeted therapies, compared to their use as single agents. The modest clinical success reached by GSIs so far could be attributed to the broad range of gamma secretase targets, resulting in toxicity and inadequate drug efficacy, thereby warranting the development of selective inhibitors directed against specific Notch pathway components. Moreover, while multiple Notch receptors are co-expressed in tumors, they do not mediate the same effects, further challenging the therapeutic efficacy of pan-Notch inhibitors. Thus, more studies are needed to understand the specific function of the different Notch receptors in cancer, as they may form distinct transcription complexes and develop specific drugs targeted to individual family members.

Acknowledgments

Author’s research is supported by grants from Italian Association for Cancer Research (AIRC-IG #2017-19923 to L.T. and AIRC-IG #2016-19032 to S.Z.) and the Fondazione Piemontese per la Ricerca sul Cancro (FPRC-ONLUS_Grant ‘5per Mille-MIUR-2013’ to L.T.). M.R. is supported by Arturo Falaschi fellowship programme at ICGEB. Thanks to Francesca Natale for editorial assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Morgan, T.H. The Physical Basis of Heredity; JB Lippincott: Philadelphia, PA, USA, 1919. [Google Scholar]
  2. Dexter, J.S. The analysis of a case of continuous variation in drosophila by a study of its linkage relations. Am. Nat. 1914, 48, 712–758. [Google Scholar] [CrossRef]
  3. Artavanis-Tsakonas, S.; Muskavitch, M.A.; Yedvobnick, B. Molecular cloning of notch, a locus affecting neurogenesis in drosophila melanogaster. Proc. Natl. Acad. Sci. USA 1983, 80, 1977–1981. [Google Scholar] [CrossRef] [PubMed]
  4. Kidd, S.; Kelley, M.R.; Young, M.W. Sequence of the notch locus of drosophila melanogaster: Relationship of the encoded protein to mammalian clotting and growth factors. Mol. Cell. Biol. 1986, 6, 3094–3108. [Google Scholar] [CrossRef] [PubMed]
  5. Aster, J.C.; Pear, W.S.; Blacklow, S.C. The varied roles of notch in cancer. Annu. Rev. Pathol. 2017, 12, 245–275. [Google Scholar] [CrossRef] [PubMed]
  6. Miele, L.; Golde, T.; Osborne, B. Notch signaling in cancer. Curr. Mol. Med. 2006, 6, 905–918. [Google Scholar] [CrossRef] [PubMed]
  7. Krebs, L.T.; Xue, Y.; Norton, C.R.; Shutter, J.R.; Maguire, M.; Sundberg, J.P.; Gallahan, D.; Closson, V.; Kitajewski, J.; Callahan, R.; et al. Notch signaling is essential for vascular morphogenesis in mice. Genes Dev. 2000, 14, 1343–1352. [Google Scholar] [PubMed]
  8. Takebe, N.; Miele, L.; Harris, P.J.; Jeong, W.; Bando, H.; Kahn, M.; Yang, S.X.; Ivy, S.P. Targeting notch, hedgehog, and wnt pathways in cancer stem cells: Clinical update. Nat. Rev. Clin. Oncol. 2015, 12, 445–464. [Google Scholar] [CrossRef] [PubMed]
  9. Koury, J.; Zhong, L.; Hao, J. Targeting signaling pathways in cancer stem cells for cancer treatment. Stem. Cells Int. 2017, 2017, 2925869. [Google Scholar] [CrossRef] [PubMed]
  10. Artavanis-Tsakonas, S.; Rand, M.D.; Lake, R.J. Notch signaling: Cell fate control and signal integration in development. Science 1999, 284, 770–776. [Google Scholar] [CrossRef] [PubMed]
  11. Del Alamo, D.; Rouault, H.; Schweisguth, F. Mechanism and significance of cis-inhibition in notch signalling. Curr. Biol. 2011, 21, R40–R47. [Google Scholar] [CrossRef] [PubMed]
  12. Bruckner, K.; Perez, L.; Clausen, H.; Cohen, S. Glycosyltransferase activity of fringe modulates notch-delta interactions. Nature 2000, 406, 411–415. [Google Scholar] [CrossRef] [PubMed]
  13. Contreras-Cornejo, H.; Saucedo-Correa, G.; Oviedo-Boyso, J.; Valdez-Alarcon, J.J.; Baizabal-Aguirre, V.M.; Cajero-Juarez, M.; Bravo-Patino, A. The csl proteins, versatile transcription factors and context dependent corepressors of the notch signaling pathway. Cell Div. 2016, 11, 12. [Google Scholar] [CrossRef] [PubMed]
  14. Andersen, P.; Uosaki, H.; Shenje, L.T.; Kwon, C. Non-canonical notch signaling: Emerging role and mechanism. Trends Cell Biol. 2012, 22, 257–265. [Google Scholar] [CrossRef] [PubMed]
  15. Ayaz, F.; Osborne, B.A. Non-canonical notch signaling in cancer and immunity. Front. Oncol. 2014, 4, 345. [Google Scholar] [CrossRef] [PubMed]
  16. Blokzijl, A.; Dahlqvist, C.; Reissmann, E.; Falk, A.; Moliner, A.; Lendahl, U.; Ibanez, C.F. Cross-talk between the notch and tgf-beta signaling pathways mediated by interaction of the notch intracellular domain with smad3. J. Cell Biol. 2003, 163, 723–728. [Google Scholar] [CrossRef] [PubMed]
  17. Liao, W.R.; Hsieh, R.H.; Hsu, K.W.; Wu, M.Z.; Tseng, M.J.; Mai, R.T.; Wu Lee, Y.H.; Yeh, T.S. The cbf1-independent notch1 signal pathway activates human c-myc expression partially via transcription factor yy1. Carcinogenesis 2007, 28, 1867–1876. [Google Scholar] [CrossRef] [PubMed]
  18. Zheng, X.; Linke, S.; Dias, J.M.; Gradin, K.; Wallis, T.P.; Hamilton, B.R.; Gustafsson, M.; Ruas, J.L.; Wilkins, S.; Bilton, R.L.; et al. Interaction with factor inhibiting hif-1 defines an additional mode of cross-coupling between the notch and hypoxia signaling pathways. Proc. Natl. Acad. Sci. USA 2008, 105, 3368–3373. [Google Scholar] [CrossRef] [PubMed]
  19. Vacca, A.; Felli, M.P.; Palermo, R.; Di Mario, G.; Calce, A.; Di Giovine, M.; Frati, L.; Gulino, A.; Screpanti, I. Notch3 and pre-tcr interaction unveils distinct nf-kappab pathways in t-cell development and leukemia. EMBO J. 2006, 25, 1000–1008. [Google Scholar] [CrossRef] [PubMed]
  20. Jin, S.; Mutvei, A.P.; Chivukula, I.V.; Andersson, E.R.; Ramskold, D.; Sandberg, R.; Lee, K.L.; Kronqvist, P.; Mamaeva, V.; Ostling, P.; et al. Non-canonical notch signaling activates il-6/jak/stat signaling in breast tumor cells and is controlled by p53 and ikkalpha/ikkbeta. Oncogene 2013, 32, 4892–4902. [Google Scholar] [CrossRef] [PubMed]
  21. Ellisen, L.W.; Bird, J.; West, D.C.; Soreng, A.L.; Reynolds, T.C.; Smith, S.D.; Sklar, J. Tan-1, the human homolog of the drosophila notch gene, is broken by chromosomal translocations in t lymphoblastic neoplasms. Cell 1991, 66, 649–661. [Google Scholar] [CrossRef]
  22. Weng, A.P.; Ferrando, A.A.; Lee, W.; Morris, J.P.T.; Silverman, L.B.; Sanchez-Irizarry, C.; Blacklow, S.C.; Look, A.T.; Aster, J.C. Activating mutations of notch1 in human t cell acute lymphoblastic leukemia. Science 2004, 306, 269–271. [Google Scholar] [CrossRef] [PubMed]
  23. Ranganathan, P.; Weaver, K.L.; Capobianco, A.J. Notch signalling in solid tumours: A little bit of everything but not all the time. Nat. Rev. Cancer 2011, 11, 338–351. [Google Scholar] [CrossRef] [PubMed]
  24. Sahlgren, C.; Gustafsson, M.V.; Jin, S.; Poellinger, L.; Lendahl, U. Notch signaling mediates hypoxia-induced tumor cell migration and invasion. Proc. Natl. Acad. Sci. USA 2008, 105, 6392–6397. [Google Scholar] [CrossRef] [PubMed]
  25. Meunier, A.; Flores, A.N.; McDermott, N.; Rivera-Figueroa, K.; Perry, A.; Lynch, T.; Redalen, K.R.; Marignol, L. Hypoxia regulates notch-3 mrna and receptor activation in prostate cancer cells. Heliyon 2016, 2, e00104. [Google Scholar] [CrossRef] [PubMed]
  26. Palomero, T.; Lim, W.K.; Odom, D.T.; Sulis, M.L.; Real, P.J.; Margolin, A.; Barnes, K.C.; O’Neil, J.; Neuberg, D.; Weng, A.P.; et al. Notch1 directly regulates c-myc and activates a feed-forward-loop transcriptional network promoting leukemic cell growth. Proc. Natl. Acad. Sci. USA 2006, 103, 18261–18266. [Google Scholar] [CrossRef] [PubMed]
  27. Zhao, K.; Wang, Q.; Wang, Y.; Huang, K.; Yang, C.; Li, Y.; Yi, K.; Kang, C. Egfr/c-myc axis regulates tgfbeta/hippo/notch pathway via epigenetic silencing mir-524 in gliomas. Cancer Lett. 2017, 406, 12–21. [Google Scholar] [CrossRef] [PubMed]
  28. Shepherd, C.; Banerjee, L.; Cheung, C.W.; Mansour, M.R.; Jenkinson, S.; Gale, R.E.; Khwaja, A. Pi3k/mtor inhibition upregulates notch-myc signalling leading to an impaired cytotoxic response. Leukemia 2013, 27, 650–660. [Google Scholar] [CrossRef] [PubMed]
  29. Sundaram, M.V. The love-hate relationship between ras and notch. Genes Dev. 2005, 19, 1825–1839. [Google Scholar] [CrossRef] [PubMed]
  30. Kefas, B.; Comeau, L.; Floyd, D.H.; Seleverstov, O.; Godlewski, J.; Schmittgen, T.; Jiang, J.; diPierro, C.G.; Li, Y.; Chiocca, E.A.; et al. The neuronal microrna mir-326 acts in a feedback loop with notch and has therapeutic potential against brain tumors. J. Neurosci. 2009, 29, 15161–15168. [Google Scholar] [CrossRef] [PubMed]
  31. Tang, Y.; Cheng, Y.S. Mir-34a inhibits pancreatic cancer progression through snail1-mediated epithelial-mesenchymal transition and the notch signaling pathway. Sci. Rep. 2017, 7, 38232. [Google Scholar] [CrossRef] [PubMed]
  32. Ji, Q.; Hao, X.; Zhang, M.; Tang, W.; Yang, M.; Li, L.; Xiang, D.; Desano, J.T.; Bommer, G.T.; Fan, D.; et al. Microrna mir-34 inhibits human pancreatic cancer tumor-initiating cells. PLoS ONE 2009, 4, e6816. [Google Scholar] [CrossRef] [PubMed]
  33. Song, G.; Zhang, Y.; Wang, L. Microrna-206 targets notch3, activates apoptosis, and inhibits tumor cell migration and focus formation. J. Biol. Chem. 2009, 284, 31921–31927. [Google Scholar] [CrossRef] [PubMed]
  34. Cohen, B.; Shimizu, M.; Izrailit, J.; Ng, N.F.; Buchman, Y.; Pan, J.G.; Dering, J.; Reedijk, M. Cyclin d1 is a direct target of jag1-mediated notch signaling in breast cancer. Breast Cancer Res. Treat. 2010, 123, 113–124. [Google Scholar] [CrossRef] [PubMed]
  35. Das, D.; Lanner, F.; Main, H.; Andersson, E.R.; Bergmann, O.; Sahlgren, C.; Heldring, N.; Hermanson, O.; Hansson, E.M.; Lendahl, U. Notch induces cyclin-d1-dependent proliferation during a specific temporal window of neural differentiation in es cells. Dev. Biol. 2010, 348, 153–166. [Google Scholar] [CrossRef] [PubMed]
  36. Mittal, S.; Subramanyam, D.; Dey, D.; Kumar, R.V.; Rangarajan, A. Cooperation of notch and ras/mapk signaling pathways in human breast carcinogenesis. Mol. Cancer 2009, 8, 128. [Google Scholar] [CrossRef] [PubMed]
  37. Meurette, O.; Stylianou, S.; Rock, R.; Collu, G.M.; Gilmore, A.P.; Brennan, K. Notch activation induces akt signaling via an autocrine loop to prevent apoptosis in breast epithelial cells. Cancer Res. 2009, 69, 5015–5022. [Google Scholar] [CrossRef] [PubMed]
  38. Li, L.; Zhao, F.; Lu, J.; Li, T.; Yang, H.; Wu, C.; Liu, Y. Notch-1 signaling promotes the malignant features of human breast cancer through nf-kappab activation. PLoS ONE 2014, 9, e95912. [Google Scholar]
  39. Gutierrez, A.; Look, A.T. Notch and pi3k-akt pathways intertwined. Cancer Cell 2007, 12, 411–413. [Google Scholar] [CrossRef] [PubMed]
  40. Kwon, O.J.; Zhang, L.; Wang, J.; Su, Q.; Feng, Q.; Zhang, X.H.; Mani, S.A.; Paulter, R.; Creighton, C.J.; Ittmann, M.M.; et al. Notch promotes tumor metastasis in a prostate-specific pten-null mouse model. J. Clin. Invest. 2016, 126, 2626–2641. [Google Scholar] [CrossRef] [PubMed]
  41. Phin, S.; Moore, M.W.; Cotter, P.D. Genomic rearrangements of pten in prostate cancer. Front. Oncol. 2013, 3, 240. [Google Scholar] [CrossRef] [PubMed]
  42. Lotan, T.L.; Carvalho, F.L.; Peskoe, S.B.; Hicks, J.L.; Good, J.; Fedor, H.; Humphreys, E.; Han, M.; Platz, E.A.; Squire, J.A.; et al. Pten loss is associated with upgrading of prostate cancer from biopsy to radical prostatectomy. Mod. Pathol. 2015, 28, 128–137. [Google Scholar] [CrossRef] [PubMed]
  43. Palomero, T.; Sulis, M.L.; Cortina, M.; Real, P.J.; Barnes, K.; Ciofani, M.; Caparros, E.; Buteau, J.; Brown, K.; Perkins, S.L.; et al. Mutational loss of pten induces resistance to notch1 inhibition in t-cell leukemia. Nat. Med. 2007, 13, 1203–1210. [Google Scholar] [CrossRef] [PubMed]
  44. Fan, L.; Wang, H.; Xia, X.; Rao, Y.; Ma, X.; Ma, D.; Wu, P.; Chen, G. Loss of e-cadherin promotes prostate cancer metastasis via upregulation of metastasis-associated gene 1 expression. Oncol. Lett. 2012, 4, 1225–1233. [Google Scholar] [CrossRef] [PubMed]
  45. Shao, S.; Zhao, X.; Zhang, X.; Luo, M.; Zuo, X.; Huang, S.; Wang, Y.; Gu, S. Notch1 signaling regulates the epithelial-mesenchymal transition and invasion of breast cancer in a slug-dependent manner. Mol. Cancer 2015, 14, 28. [Google Scholar] [CrossRef] [PubMed]
  46. Fender, A.W.; Nutter, J.M.; Fitzgerald, T.L.; Bertrand, F.E.; Sigounas, G. Notch-1 promotes stemness and epithelial to mesenchymal transition in colorectal cancer. J. Cell Biochem. 2015, 116, 2517–2527. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, Z.; Li, Y.; Kong, D.; Sarkar, F.H. The role of notch signaling pathway in epithelial-mesenchymal transition (emt) during development and tumor aggressiveness. Curr. Drug Targets 2010, 11, 745–751. [Google Scholar] [CrossRef] [PubMed]
  48. Rehman, M.; Gurrapu, S.; Cagnoni, G.; Capparuccia, L.; Tamagnone, L. Plexind1 is a novel transcriptional target and effector of notch signaling in cancer cells. PLoS ONE 2016, 11, e0164660. [Google Scholar] [CrossRef] [PubMed]
  49. Casazza, A.; Finisguerra, V.; Capparuccia, L.; Camperi, A.; Swiercz, J.M.; Rizzolio, S.; Rolny, C.; Christensen, C.; Bertotti, A.; Sarotto, I.; et al. Sema3e-plexin d1 signaling drives human cancer cell invasiveness and metastatic spreading in mice. J. Clin. Invest. 2010, 120, 2684–2698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Tseng, C.H.; Murray, K.D.; Jou, M.F.; Hsu, S.M.; Cheng, H.J.; Huang, P.H. Sema3e/plexin-d1 mediated epithelial-to-mesenchymal transition in ovarian endometrioid cancer. PLoS ONE 2011, 6, e19396. [Google Scholar] [CrossRef] [PubMed]
  51. Kim, J.; Oh, W.J.; Gaiano, N.; Yoshida, Y.; Gu, C. Semaphorin 3e-plexin-d1 signaling regulates vegf function in developmental angiogenesis via a feedback mechanism. Genes Dev. 2011, 25, 1399–1411. [Google Scholar] [CrossRef] [PubMed]
  52. Avila, J.L.; Kissil, J.L. Notch signaling in pancreatic cancer: Oncogene or tumor suppressor? Trends Mol. Med. 2013, 19, 320–327. [Google Scholar] [CrossRef] [PubMed]
  53. Yap, L.F.; Lee, D.; Khairuddin, A.; Pairan, M.F.; Puspita, B.; Siar, C.H.; Paterson, I.C. The opposing roles of notch signalling in head and neck cancer: A mini review. Oral. Dis. 2015, 21, 850–857. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, N.J.; Sanborn, Z.; Arnett, K.L.; Bayston, L.J.; Liao, W.; Proby, C.M.; Leigh, I.M.; Collisson, E.A.; Gordon, P.B.; Jakkula, L.; et al. Loss-of-function mutations in notch receptors in cutaneous and lung squamous cell carcinoma. Proc. Natl. Acad. Sci. USA 2011, 108, 17761–17766. [Google Scholar] [CrossRef] [PubMed]
  55. Rangarajan, A.; Talora, C.; Okuyama, R.; Nicolas, M.; Mammucari, C.; Oh, H.; Aster, J.C.; Krishna, S.; Metzger, D.; Chambon, P.; et al. Notch signaling is a direct determinant of keratinocyte growth arrest and entry into differentiation. EMBO J. 2001, 20, 3427–3436. [Google Scholar] [CrossRef] [PubMed]
  56. Rampias, T.; Vgenopoulou, P.; Avgeris, M.; Polyzos, A.; Stravodimos, K.; Valavanis, C.; Scorilas, A.; Klinakis, A. A new tumor suppressor role for the notch pathway in bladder cancer. Nat. Med. 2014, 20, 1199–1205. [Google Scholar] [CrossRef] [PubMed]
  57. Hu, Y.; Su, H.; Li, X.; Guo, G.; Cheng, L.; Qin, R.; Qing, G.; Liu, H. The notch ligand jagged2 promotes pancreatic cancer metastasis independent of notch signaling activation. Mol. Cancer Ther. 2015, 14, 289–297. [Google Scholar] [CrossRef] [PubMed]
  58. Ortica, S.; Tarantino, N.; Aulner, N.; Israel, A.; Gupta-Rossi, N. The 4 notch receptors play distinct and antagonistic roles in the proliferation and hepatocytic differentiation of liver progenitors. FASEB J. 2014, 28, 603–614. [Google Scholar] [CrossRef] [PubMed]
  59. Chu, D.; Zhang, Z.; Zhou, Y.; Wang, W.; Li, Y.; Zhang, H.; Dong, G.; Zhao, Q.; Ji, G. Notch1 and notch2 have opposite prognostic effects on patients with colorectal cancer. Ann. Oncol. 2011, 22, 2440–2447. [Google Scholar] [CrossRef] [PubMed]
  60. Chu, D.; Zheng, J.; Wang, W.; Zhao, Q.; Li, Y.; Li, J.; Xie, H.; Zhang, H.; Dong, G.; Xu, C.; et al. Notch2 expression is decreased in colorectal cancer and related to tumor differentiation status. Ann. Surg. Oncol. 2009, 16, 3259–3266. [Google Scholar] [CrossRef] [PubMed]
  61. Choy, L.; Hagenbeek, T.J.; Solon, M.; French, D.; Finkle, D.; Shelton, A.; Venook, R.; Brauer, M.J.; Siebel, C.W. Constitutive notch3 signaling promotes the growth of basal breast cancers. Cancer Res. 2017, 77, 1439–1452. [Google Scholar] [CrossRef] [PubMed]
  62. Danza, G.; Di Serio, C.; Ambrosio, M.R.; Sturli, N.; Lonetto, G.; Rosati, F.; Rocca, B.J.; Ventimiglia, G.; del Vecchio, M.T.; Prudovsky, I.; et al. Notch3 is activated by chronic hypoxia and contributes to the progression of human prostate cancer. Int. J. Cancer 2013, 133, 2577–2586. [Google Scholar] [CrossRef] [PubMed]
  63. Zhou, L.; Zhang, N.; Song, W.; You, N.; Li, Q.; Sun, W.; Zhang, Y.; Wang, D.; Dou, K. The significance of notch1 compared with notch3 in high metastasis and poor overall survival in hepatocellular carcinoma. PLoS ONE 2013, 8, e57382. [Google Scholar] [CrossRef] [PubMed]
  64. Kayamori, K.; Katsube, K.; Sakamoto, K.; Ohyama, Y.; Hirai, H.; Yukimori, A.; Ohata, Y.; Akashi, T.; Saitoh, M.; Harada, K.; et al. Notch3 is induced in cancer-associated fibroblasts and promotes angiogenesis in oral squamous cell carcinoma. PLoS ONE 2016, 11, e0154112. [Google Scholar] [CrossRef] [PubMed]
  65. Bernasconi-Elias, P.; Hu, T.; Jenkins, D.; Firestone, B.; Gans, S.; Kurth, E.; Capodieci, P.; Deplazes-Lauber, J.; Petropoulos, K.; Thiel, P.; et al. Characterization of activating mutations of notch3 in t-cell acute lymphoblastic leukemia and anti-leukemic activity of notch3 inhibitory antibodies. Oncogene 2016, 35, 6077–6086. [Google Scholar] [CrossRef] [PubMed]
  66. Choi, S.H.; Severson, E.; Pear, W.S.; Liu, X.S.; Aster, J.C.; Blacklow, S.C. The common oncogenomic program of notch1 and notch3 signaling in t-cell acute lymphoblastic leukemia. PLoS ONE 2017, 12, e0185762. [Google Scholar] [CrossRef] [PubMed]
  67. Cui, H.; Kong, Y.; Xu, M.; Zhang, H. Notch3 functions as a tumor suppressor by controlling cellular senescence. Cancer Res. 2013, 73, 3451–3459. [Google Scholar] [CrossRef] [PubMed]
  68. Zhang, X.; Liu, X.; Luo, J.; Xiao, W.; Ye, X.; Chen, M.; Li, Y.; Zhang, G.J. Notch3 inhibits epithelial-mesenchymal transition by activating kibra-mediated hippo/yap signaling in breast cancer epithelial cells. Oncogenesis 2016, 5, e269. [Google Scholar] [CrossRef] [PubMed]
  69. Nagamatsu, I.; Onishi, H.; Matsushita, S.; Kubo, M.; Kai, M.; Imaizumi, A.; Nakano, K.; Hattori, M.; Oda, Y.; Tanaka, M.; et al. Notch4 is a potential therapeutic target for triple-negative breast cancer. Anticancer Res. 2014, 34, 69–80. [Google Scholar] [PubMed]
  70. Bui, Q.T.; Im, J.H.; Jeong, S.B.; Kim, Y.M.; Lim, S.C.; Kim, B.; Kang, K.W. Essential role of notch4/stat3 signaling in epithelial-mesenchymal transition of tamoxifen-resistant human breast cancer. Cancer Lett. 2017, 390, 115–125. [Google Scholar] [CrossRef] [PubMed]
  71. Qian, C.J.; Chen, Y.Y.; Zhang, X.; Liu, F.Q.; Yue, T.T.; Ye, B.; Yao, J. Notch4 inhibition reduces migration and invasion and enhances sensitivity to docetaxel by inhibiting akt/fascin in pancreatic cancer cells. Oncol. Lett. 2016, 12, 3499–3505. [Google Scholar] [CrossRef] [PubMed]
  72. Bonyadi Rad, E.; Hammerlindl, H.; Wels, C.; Popper, U.; Ravindran Menon, D.; Breiteneder, H.; Kitzwoegerer, M.; Hafner, C.; Herlyn, M.; Bergler, H.; et al. Notch4 signaling induces a mesenchymal-epithelial-like transition in melanoma cells to suppress malignant behaviors. Cancer Res. 2016, 76, 1690–1697. [Google Scholar] [CrossRef] [PubMed]
  73. Kulic, I.; Robertson, G.; Chang, L.; Baker, J.H.; Lockwood, W.W.; Mok, W.; Fuller, M.; Fournier, M.; Wong, N.; Chou, V.; et al. Loss of the notch effector rbpj promotes tumorigenesis. J. Exp. Med. 2015, 212, 37–52. [Google Scholar] [CrossRef] [PubMed]
  74. Yu, K.; Ganesan, K.; Tan, L.K.; Laban, M.; Wu, J.; Zhao, X.D.; Li, H.; Leung, C.H.; Zhu, Y.; Wei, C.L.; et al. A precisely regulated gene expression cassette potently modulates metastasis and survival in multiple solid cancers. PLoS Genet. 2008, 4, e1000129. [Google Scholar] [CrossRef] [PubMed]
  75. Xie, Q.; Wu, Q.; Kim, L.; Miller, T.E.; Liau, B.B.; Mack, S.C.; Yang, K.; Factor, D.C.; Fang, X.; Huang, Z.; et al. Rbpj maintains brain tumor-initiating cells through cdk9-mediated transcriptional elongation. J. Clin. Invest. 2016, 126, 2757–2772. [Google Scholar] [CrossRef] [PubMed]
  76. De Kloe, G.E.; De Strooper, B. Small molecules that inhibit notch signaling. Methods Mol. Biol. 2014, 1187, 311–322. [Google Scholar] [PubMed]
  77. Liebler, S.S.; Feldner, A.; Adam, M.G.; Korff, T.; Augustin, H.G.; Fischer, A. No evidence for a functional role of bi-directional notch signaling during angiogenesis. PLoS ONE 2012, 7, e53074. [Google Scholar] [CrossRef] [PubMed]
  78. Vidal, G.A.; Naresh, A.; Marrero, L.; Jones, F.E. Presenilin-dependent gamma-secretase processing regulates multiple erbb4/her4 activities. J. Biol. Chem. 2005, 280, 19777–19783. [Google Scholar] [CrossRef] [PubMed]
  79. Murakami, D.; Okamoto, I.; Nagano, O.; Kawano, Y.; Tomita, T.; Iwatsubo, T.; De Strooper, B.; Yumoto, E.; Saya, H. Presenilin-dependent gamma-secretase activity mediates the intramembranous cleavage of cd44. Oncogene 2003, 22, 1511–1516. [Google Scholar] [CrossRef] [PubMed]
  80. Olson, R.E.; Albright, C.F. Recent progress in the medicinal chemistry of gamma-secretase inhibitors. Curr. Top. Med. Chem. 2008, 8, 17–33. [Google Scholar] [CrossRef] [PubMed]
  81. Dovey, H.F.; John, V.; Anderson, J.P.; Chen, L.Z.; de Saint Andrieu, P.; Fang, L.Y.; Freedman, S.B.; Folmer, B.; Goldbach, E.; Holsztynska, E.J.; et al. Functional gamma-secretase inhibitors reduce beta-amyloid peptide levels in brain. J. Neurochem. 2001, 76, 173–181. [Google Scholar] [CrossRef] [PubMed]
  82. Saito, N.; Fu, J.; Zheng, S.; Yao, J.; Wang, S.; Liu, D.D.; Yuan, Y.; Sulman, E.P.; Lang, F.F.; Colman, H.; et al. A high notch pathway activation predicts response to gamma secretase inhibitors in proneural subtype of glioma tumor-initiating cells. Stem Cells 2014, 32, 301–312. [Google Scholar] [CrossRef] [PubMed]
  83. Gavai, A.V.; Quesnelle, C.; Norris, D.; Han, W.C.; Gill, P.; Shan, W.; Balog, A.; Chen, K.; Tebben, A.; Rampulla, R.; et al. Discovery of clinical candidate bms-906024: A potent pan-notch inhibitor for the treatment of leukemia and solid tumors. ACS Med. Chem. Lett. 2015, 6, 523–527. [Google Scholar] [CrossRef] [PubMed]
  84. Bolos, V.; Mira, E.; Martinez-Poveda, B.; Luxan, G.; Canamero, M.; Martinez, A.C.; Manes, S.; de la Pompa, J.L. Notch activation stimulates migration of breast cancer cells and promotes tumor growth. Breast Cancer Res. 2013, 15, R54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Su, F.; Zhu, S.; Ruan, J.; Muftuoglu, Y.; Zhang, L.; Yuan, Q. Combination therapy of ry10-4 with the gamma-secretase inhibitor dapt shows promise in treating her2-amplified breast cancer. Oncotarget 2016, 7, 4142–4154. [Google Scholar] [PubMed]
  86. Fan, X.; Khaki, L.; Zhu, T.S.; Soules, M.E.; Talsma, C.E.; Gul, N.; Koh, C.; Zhang, J.; Li, Y.M.; Maciaczyk, J.; et al. Notch pathway blockade depletes cd133-positive glioblastoma cells and inhibits growth of tumor neurospheres and xenografts. Stem Cells 2010, 28, 5–16. [Google Scholar] [CrossRef] [PubMed]
  87. Strosberg, J.R.; Yeatman, T.; Weber, J.; Coppola, D.; Schell, M.J.; Han, G.; Almhanna, K.; Kim, R.; Valone, T.; Jump, H.; et al. A phase ii study of ro4929097 in metastatic colorectal cancer. Eur. J. Cancer 2012, 48, 997–1003. [Google Scholar] [CrossRef] [PubMed]
  88. Samon, J.B.; Castillo-Martin, M.; Hadler, M.; Ambesi-Impiobato, A.; Paietta, E.; Racevskis, J.; Wiernik, P.H.; Rowe, J.M.; Jakubczak, J.; Randolph, S.; et al. Preclinical analysis of the gamma-secretase inhibitor pf-03084014 in combination with glucocorticoids in t-cell acute lymphoblastic leukemia. Mol. Cancer Ther. 2012, 11, 1565–1575. [Google Scholar] [CrossRef] [PubMed]
  89. De Jesus-Acosta, A.; Laheru, D.; Maitra, A.; Arcaroli, J.; Rudek, M.A.; Dasari, A.; Blatchford, P.J.; Quackenbush, K.; Messersmith, W. A phase ii study of the gamma secretase inhibitor ro4929097 in patients with previously treated metastatic pancreatic adenocarcinoma. Invest. New Drugs 2014, 32, 739–745. [Google Scholar] [CrossRef] [PubMed]
  90. Krop, I.; Demuth, T.; Guthrie, T.; Wen, P.Y.; Mason, W.P.; Chinnaiyan, P.; Butowski, N.; Groves, M.D.; Kesari, S.; Freedman, S.J.; et al. Phase i pharmacologic and pharmacodynamic study of the gamma secretase (notch) inhibitor mk-0752 in adult patients with advanced solid tumors. J. Clin. Oncol. 2012, 30, 2307–2313. [Google Scholar] [CrossRef] [PubMed]
  91. O’Neil, J.; Grim, J.; Strack, P.; Rao, S.; Tibbitts, D.; Winter, C.; Hardwick, J.; Welcker, M.; Meijerink, J.P.; Pieters, R.; et al. Fbw7 mutations in leukemic cells mediate notch pathway activation and resistance to gamma-secretase inhibitors. J. Exp. Med. 2007, 204, 1813–1824. [Google Scholar] [CrossRef] [PubMed]
  92. Yashiro-Ohtani, Y.; Wang, H.; Zang, C.; Arnett, K.L.; Bailis, W.; Ho, Y.; Knoechel, B.; Lanauze, C.; Louis, L.; Forsyth, K.S.; et al. Long-range enhancer activity determines myc sensitivity to notch inhibitors in t cell leukemia. Proc. Natl. Acad. Sci. USA 2014, 111, E4946–E4953. [Google Scholar] [CrossRef] [PubMed]
  93. Doody, R.S.; Raman, R.; Sperling, R.A.; Seimers, E.; Sethuraman, G.; Mohs, R.; Farlow, M.; Iwatsubo, T.; Vellas, B.; Sun, X.; et al. Peripheral and central effects of gamma-secretase inhibition by semagacestat in alzheimer’s disease. Alzheimers Res. Ther. 2015, 7, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Doody, R.S.; Raman, R.; Farlow, M.; Iwatsubo, T.; Vellas, B.; Joffe, S.; Kieburtz, K.; He, F.; Sun, X.; Thomas, R.G.; et al. A phase 3 trial of semagacestat for treatment of alzheimer’s disease. N. Eng. J. Med. 2013, 369, 341–350. [Google Scholar] [CrossRef] [PubMed]
  95. Ryeom, S.W. The cautionary tale of side effects of chronic notch1 inhibition. J. Clin. Invest. 2011, 121, 508–509. [Google Scholar] [CrossRef] [PubMed]
  96. Liu, Z.; Turkoz, A.; Jackson, E.N.; Corbo, J.C.; Engelbach, J.A.; Garbow, J.R.; Piwnica-Worms, D.R.; Kopan, R. Notch1 loss of heterozygosity causes vascular tumors and lethal hemorrhage in mice. J. Clin. Invest. 2011, 121, 800–808. [Google Scholar] [CrossRef] [PubMed]
  97. VanDussen, K.L.; Carulli, A.J.; Keeley, T.M.; Patel, S.R.; Puthoff, B.J.; Magness, S.T.; Tran, I.T.; Maillard, I.; Siebel, C.; Kolterud, A.; et al. Notch signaling modulates proliferation and differentiation of intestinal crypt base columnar stem cells. Development 2012, 139, 488–497. [Google Scholar] [CrossRef] [PubMed]
  98. Astudillo, L.; Da Silva, T.G.; Wang, Z.; Han, X.; Jin, K.; VanWye, J.; Zhu, X.; Weaver, K.; Oashi, T.; Lopes, P.E.; et al. The small molecule imr-1 inhibits the notch transcriptional activation complex to suppress tumorigenesis. Cancer Res. 2016, 76, 3593–3603. [Google Scholar] [CrossRef] [PubMed]
  99. Kamstrup, M.R.; Biskup, E.; Manfe, V.; Savorani, C.; Liszewski, W.; Wiren, J.; Specht, L.; Gniadecki, R. Chemotherapeutic treatment is associated with notch1 induction in cutaneous t-cell lymphoma. Leuk. Lymphoma 2017, 58, 171–178. [Google Scholar] [CrossRef] [PubMed]
  100. Gu, F.; Ma, Y.; Zhang, Z.; Zhao, J.; Kobayashi, H.; Zhang, L.; Fu, L. Expression of stat3 and notch1 is associated with cisplatin resistance in head and neck squamous cell carcinoma. Oncol. Rep. 2010, 23, 671–676. [Google Scholar] [PubMed]
  101. Meng, R.D.; Shelton, C.C.; Li, Y.M.; Qin, L.X.; Notterman, D.; Paty, P.B.; Schwartz, G.K. Gamma-secretase inhibitors abrogate oxaliplatin-induced activation of the notch-1 signaling pathway in colon cancer cells resulting in enhanced chemosensitivity. Cancer Res. 2009, 69, 573–582. [Google Scholar] [CrossRef] [PubMed]
  102. Yao, J.; Qian, C. Inhibition of notch3 enhances sensitivity to gemcitabine in pancreatic cancer through an inactivation of pi3k/akt-dependent pathway. Med. Oncol. 2010, 27, 1017–1022. [Google Scholar] [CrossRef] [PubMed]
  103. Kim, B.; Stephen, S.L.; Hanby, A.M.; Horgan, K.; Perry, S.L.; Richardson, J.; Roundhill, E.A.; Valleley, E.M.; Verghese, E.T.; Williams, B.J.; et al. Chemotherapy induces notch1-dependent mrp1 up-regulation, inhibition of which sensitizes breast cancer cells to chemotherapy. BMC Cancer 2015, 15, 634. [Google Scholar] [CrossRef] [PubMed]
  104. Li, Z.L.; Chen, C.; Yang, Y.; Wang, C.; Yang, T.; Yang, X.; Liu, S.C. Gamma secretase inhibitor enhances sensitivity to doxorubicin in mda-mb-231 cells. Int. J. Clin. Exp. Pathol. 2015, 8, 4378–4387. [Google Scholar] [PubMed]
  105. Dong, Y.; Li, A.; Wang, J.; Weber, J.D.; Michel, L.S. Synthetic lethality through combined notch-epidermal growth factor receptor pathway inhibition in basal-like breast cancer. Cancer Res. 2010, 70, 5465–5474. [Google Scholar] [CrossRef] [PubMed]
  106. Pandya, K.; Meeke, K.; Clementz, A.G.; Rogowski, A.; Roberts, J.; Miele, L.; Albain, K.S.; Osipo, C. Targeting both notch and erbb-2 signalling pathways is required for prevention of erbb-2-positive breast tumour recurrence. Br. J. Cancer 2011, 105, 796–806. [Google Scholar] [CrossRef] [PubMed]
  107. Cui, D.; Dai, J.; Keller, J.M.; Mizokami, A.; Xia, S.; Keller, E.T. Notch pathway inhibition using pf-03084014, a gamma-secretase inhibitor (gsi), enhances the antitumor effect of docetaxel in prostate cancer. Clin. Cancer Res. 2015, 21, 4619–4629. [Google Scholar] [CrossRef] [PubMed]
  108. Jin, R.; Nakada, M.; Teng, L.; Furuta, T.; Sabit, H.; Hayashi, Y.; Demuth, T.; Hirao, A.; Sato, H.; Zhao, G.; et al. Combination therapy using notch and akt inhibitors is effective for suppressing invasion but not proliferation in glioma cells. Neurosci. Lett. 2013, 534, 316–321. [Google Scholar] [CrossRef] [PubMed]
  109. Ambrogio, C.; Gomez-Lopez, G.; Falcone, M.; Vidal, A.; Nadal, E.; Crosetto, N.; Blasco, R.B.; Fernandez-Marcos, P.J.; Sanchez-Cespedes, M.; Ren, X.; et al. Combined inhibition of ddr1 and notch signaling is a therapeutic strategy for kras-driven lung adenocarcinoma. Nat. Med. 2016, 22, 270–277. [Google Scholar] [CrossRef] [PubMed]
  110. Li, M.; Chen, F.; Clifton, N.; Sullivan, D.M.; Dalton, W.S.; Gabrilovich, D.I.; Nefedova, Y. Combined inhibition of notch signaling and bcl-2/bcl-xl results in synergistic antimyeloma effect. Mol. Cancer Ther. 2010, 9, 3200–3209. [Google Scholar] [CrossRef] [PubMed]
  111. Weng, A.P.; Millholland, J.M.; Yashiro-Ohtani, Y.; Arcangeli, M.L.; Lau, A.; Wai, C.; Del Bianco, C.; Rodriguez, C.G.; Sai, H.; Tobias, J.; et al. C-myc is an important direct target of notch1 in t-cell acute lymphoblastic leukemia/lymphoma. Genes Dev. 2006, 20, 2096–2109. [Google Scholar] [CrossRef] [PubMed]
  112. Falk, R.; Falk, A.; Dyson, M.R.; Melidoni, A.N.; Parthiban, K.; Young, J.L.; Roake, W.; McCafferty, J. Generation of anti-notch antibodies and their application in blocking notch signalling in neural stem cells. Methods 2012, 58, 69–78. [Google Scholar] [CrossRef] [PubMed]
  113. Yen, W.C.; Fischer, M.M.; Axelrod, F.; Bond, C.; Cain, J.; Cancilla, B.; Henner, W.R.; Meisner, R.; Sato, A.; Shah, J.; et al. Targeting notch signaling with a notch2/notch3 antagonist (tarextumab) inhibits tumor growth and decreases tumor-initiating cell frequency. Clin. Cancer Res. 2015, 21, 2084–2095. [Google Scholar] [CrossRef] [PubMed]
  114. O’Reilly, E.M.; Smith, L.S.; Bendell, J.C.; Rangwala, F.A.; Schmidt, W.; Stephenson, J.; Kapoun, A.; Xu, L.; Hill, D.; Zhou, L.; et al. Phase ib of anticancer stem cell antibody omp-59r5 (anti-notch2/3) in combination with nab-paclitaxel and gemcitabine (nab-p+gem) in patients (pts) with untreated metastatic pancreatic cancer (mpc). J. Clin. Oncol. 2014, 32, 220. [Google Scholar] [CrossRef]
  115. Wu, Y.; Cain-Hom, C.; Choy, L.; Hagenbeek, T.J.; de Leon, G.P.; Chen, Y.; Finkle, D.; Venook, R.; Wu, X.; Ridgway, J.; et al. Therapeutic antibody targeting of individual notch receptors. Nature 2010, 464, 1052–1057. [Google Scholar] [CrossRef] [PubMed]
  116. Brunner, A.; Cattaruzza, F.; Yen, W.-C.; Yeung, P.; Fischer, M.; Cancilla, B.; O’Young, G.; Tam, R.; Liu, Y.-W.; Gurney, A.; et al. Abstract 4652: Effects of anti-dll4 treatment on non-small cell lung cancer (nsclc) human xenograft tumors. Cancer Res. 2016, 76, 4652. [Google Scholar] [CrossRef]
  117. Ridgway, J.; Zhang, G.; Wu, Y.; Stawicki, S.; Liang, W.C.; Chanthery, Y.; Kowalski, J.; Watts, R.J.; Callahan, C.; Kasman, I.; et al. Inhibition of dll4 signalling inhibits tumour growth by deregulating angiogenesis. Nature 2006, 444, 1083–1087. [Google Scholar] [CrossRef] [PubMed]
  118. Huang, J.; Hu, W.; Hu, L.; Previs, R.A.; Dalton, H.J.; Yang, X.Y.; Sun, Y.; McGuire, M.; Rupaimoole, R.; Nagaraja, A.S.; et al. Dll4 inhibition plus aflibercept markedly reduces ovarian tumor growth. Mol. Cancer Ther. 2016, 15, 1344–1352. [Google Scholar] [CrossRef] [PubMed]
  119. Funahashi, Y.; Hernandez, S.L.; Das, I.; Ahn, A.; Huang, J.; Vorontchikhina, M.; Sharma, A.; Kanamaru, E.; Borisenko, V.; Desilva, D.M.; et al. A notch1 ectodomain construct inhibits endothelial notch signaling, tumor growth, and angiogenesis. Cancer Res. 2008, 68, 4727–4735. [Google Scholar] [CrossRef] [PubMed]
  120. Kuramoto, T.; Goto, H.; Mitsuhashi, A.; Tabata, S.; Ogawa, H.; Uehara, H.; Saijo, A.; Kakiuchi, S.; Maekawa, Y.; Yasutomo, K.; et al. Dll4-fc, an inhibitor of dll4-notch signaling, suppresses liver metastasis of small cell lung cancer cells through the downregulation of the nf-kappab activity. Mol. Cancer Ther. 2012, 11, 2578–2587. [Google Scholar] [CrossRef] [PubMed]
  121. Baladron, V.; Ruiz-Hidalgo, M.J.; Nueda, M.L.; Diaz-Guerra, M.J.; Garcia-Ramirez, J.J.; Bonvini, E.; Gubina, E.; Laborda, J. Dlk acts as a negative regulator of notch1 activation through interactions with specific egf-like repeats. Exp. Cell Res. 2005, 303, 343–359. [Google Scholar] [CrossRef] [PubMed]
  122. Klose, R.; Berger, C.; Moll, I.; Adam, M.G.; Schwarz, F.; Mohr, K.; Augustin, H.G.; Fischer, A. Soluble notch ligand and receptor peptides act antagonistically during angiogenesis. Cardiova. Res. 2015, 107, 153–163. [Google Scholar] [CrossRef] [PubMed]
  123. Kangsamaksin, T.; Murtomaki, A.; Kofler, N.M.; Cuervo, H.; Chaudhri, R.A.; Tattersall, I.W.; Rosenstiel, P.E.; Shawber, C.J.; Kitajewski, J. Notch decoys that selectively block dll/notch or jag/notch disrupt angiogenesis by unique mechanisms to inhibit tumor growth. Cancer Discov. 2015, 5, 182–197. [Google Scholar] [CrossRef] [PubMed]
  124. Moellering, R.E.; Cornejo, M.; Davis, T.N.; Del Bianco, C.; Aster, J.C.; Blacklow, S.C.; Kung, A.L.; Gilliland, D.G.; Verdine, G.L.; Bradner, J.E. Direct inhibition of the notch transcription factor complex. Nature 2009, 462, 182–188. [Google Scholar] [CrossRef] [PubMed]
  125. Lundin, K.E.; Gissberg, O.; Smith, C.I. Oligonucleotide therapies: The past and the present. Hum. Gene Ther. 2015, 26, 475–485. [Google Scholar] [CrossRef] [PubMed]
  126. Nakazawa, M.; Ishii, H.; Aono, H.; Takai, M.; Honda, T.; Aratani, S.; Fukamizu, A.; Nakamura, H.; Yoshino, S.; Kobata, T.; et al. Role of notch-1 intracellular domain in activation of rheumatoid synoviocytes. Arthritis Rheum. 2001, 44, 1545–1554. [Google Scholar] [CrossRef]
  127. Shan, L.; Aster, J.C.; Sklar, J.; Sunday, M.E. Notch-1 regulates pulmonary neuroendocrine cell differentiation in cell lines and in transgenic mice. Am. J. Physiol. Lung Cell. Mol. Physiol. 2007, 292, L500–L509. [Google Scholar] [CrossRef] [PubMed]
  128. Zimrin, A.B.; Pepper, M.S.; McMahon, G.A.; Nguyen, F.; Montesano, R.; Maciag, T. An antisense oligonucleotide to the notch ligand jagged enhances fibroblast growth factor-induced angiogenesis in vitro. J. Biol. Chem. 1996, 271, 32499–32502. [Google Scholar] [CrossRef] [PubMed]
  129. Wang, Y.; Chan, S.L.; Miele, L.; Yao, P.J.; Mackes, J.; Ingram, D.K.; Mattson, M.P.; Furukawa, K. Involvement of notch signaling in hippocampal synaptic plasticity. Proc. Natl. Acad. Sci. USA 2004, 101, 9458–9462. [Google Scholar] [CrossRef] [PubMed]
  130. Souilhol, C.; Cormier, S.; Monet, M.; Vandormael-Pournin, S.; Joutel, A.; Babinet, C.; Cohen-Tannoudji, M. Nas transgenic mouse line allows visualization of notch pathway activity in vivo. Genesis 2006, 44, 277–286. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Therapeutic targets in the Notch signaling pathway. The Notch pathway can be targeted by small molecule Gamma Secretase Inhibitors (GSI), antibodies (Anti-Notch, Anti-Dll4), decoys, and peptides.
Figure 1. Therapeutic targets in the Notch signaling pathway. The Notch pathway can be targeted by small molecule Gamma Secretase Inhibitors (GSI), antibodies (Anti-Notch, Anti-Dll4), decoys, and peptides.
Molecules 23 00431 g001
Figure 2. GSI structures. Structures of some commonly used GSI, which are used in vitro and in clinical trials.
Figure 2. GSI structures. Structures of some commonly used GSI, which are used in vitro and in clinical trials.
Molecules 23 00431 g002
Figure 3. Functional effects of gamma secretase inhibitors. The pie chart represents the main biological activities promoted or inhibited by GSI, based on the prevalence of their report in literature (approx. 450 Pubmed-indexed research papers related to the use of GSI were individually analyzed).
Figure 3. Functional effects of gamma secretase inhibitors. The pie chart represents the main biological activities promoted or inhibited by GSI, based on the prevalence of their report in literature (approx. 450 Pubmed-indexed research papers related to the use of GSI were individually analyzed).
Molecules 23 00431 g003
Figure 4. Clinical trials (named by the NCT ClinicalTrials.gov identifier) for the treatment of the indicated tumor types based on the use of Notch inhibitors, either as a single agents (upper panel) or in combination with other drugs (lower panel).
Figure 4. Clinical trials (named by the NCT ClinicalTrials.gov identifier) for the treatment of the indicated tumor types based on the use of Notch inhibitors, either as a single agents (upper panel) or in combination with other drugs (lower panel).
Molecules 23 00431 g004

Share and Cite

MDPI and ACS Style

Tamagnone, L.; Zacchigna, S.; Rehman, M. Taming the Notch Transcriptional Regulator for Cancer Therapy. Molecules 2018, 23, 431. https://doi.org/10.3390/molecules23020431

AMA Style

Tamagnone L, Zacchigna S, Rehman M. Taming the Notch Transcriptional Regulator for Cancer Therapy. Molecules. 2018; 23(2):431. https://doi.org/10.3390/molecules23020431

Chicago/Turabian Style

Tamagnone, Luca, Serena Zacchigna, and Michael Rehman. 2018. "Taming the Notch Transcriptional Regulator for Cancer Therapy" Molecules 23, no. 2: 431. https://doi.org/10.3390/molecules23020431

Article Metrics

Back to TopTop