Next Article in Journal
Negative Survival Impact of Occult Lymph Node Involvement in Small HER2-Positive Early Breast Cancer Treated by Up-Front Surgery
Next Article in Special Issue
Pro-Apoptotic Activity of MCL-1 Inhibitor in Trametinib-Resistant Melanoma Cells Depends on Their Phenotypes and Is Modulated by Reversible Alterations Induced by Trametinib Withdrawal
Previous Article in Journal
Recent Insights into the Biomarkers, Molecular Targets and Mechanisms of Non-Alcoholic Steatohepatitis-Driven Hepatocarcinogenesis
Previous Article in Special Issue
Potential Role of Carbon Nanomaterials in the Treatment of Malignant Brain Gliomas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modulation of Notch Signaling by Small-Molecular Compounds and Its Potential in Anticancer Studies

Department of Biochemistry and Molecular Biology, Medical University of Lublin, 20-093 Lublin, Poland
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(18), 4563; https://doi.org/10.3390/cancers15184563
Submission received: 1 August 2023 / Revised: 3 September 2023 / Accepted: 7 September 2023 / Published: 14 September 2023
(This article belongs to the Special Issue Molecular Mechanisms of Anti-cancer Drugs Action)

Abstract

:

Simple Summary

The dysregulation of Notch signaling plays an important role in cancer development. Numerous attempts are made to modulate Notch signaling for positive therapeutic effects. For this purpose, a broad group of small-molecular compounds are used. These small compounds have several features desirable in laboratory and clinical practice, such as low price, easy availability, and the possibility of standardizing the dose, and will be presented in this article.

Abstract

Notch signaling is responsible for conveying messages between cells through direct contact, playing a pivotal role in tissue development and homeostasis. The modulation of Notch-related processes, such as cell growth, differentiation, viability, and cell fate, offer opportunities to better understand and prevent disease progression, including cancer. Currently, research efforts are mainly focused on attempts to inhibit Notch signaling in tumors with strong oncogenic, gain-of-function (GoF) or hyperactivation of Notch signaling. The goal is to reduce the growth and proliferation of cancer cells, interfere with neo-angiogenesis, increase chemosensitivity, potentially target cancer stem cells, tumor dormancy, and invasion, and induce apoptosis. Attempts to pharmacologically enhance or restore disturbed Notch signaling for anticancer therapies are less frequent. However, in some cancer types, such as squamous cell carcinomas, preferentially, loss-of-function (LoF) mutations have been confirmed, and restoring but not blocking Notch functions may be beneficial for therapy. The modulation of Notch signaling can be performed at several key levels related to NOTCH receptor expression, translation, posttranslational (proteolytic) processing, glycosylation, transport, and activation. This further includes blocking the interaction with Notch-related nuclear DNA transcription. Examples of small-molecular chemical compounds, that modulate individual elements of Notch signaling at the mentioned levels, have been described in the recent literature.

1. Canonical and Non-Canonical Notch Pathway

Notch signaling (Figure 1) in mammals is based on the activation of one of the four single membrane pass NOTCH mechanoreceptors (NOTCH1, 2, 3, and 4) by their transmembrane ligands, Delta-like-1,3,4 (DLL1; 3; and 4) or Jagged 1, 2 (JAG1 and JAG2), collectively known as Delta/Serrate/Lag-2 (DSL) ligands. The basic structure of the NOTCH receptor can be distinguished as (a) the NOTCH extracellular domain (NECD) with the region containing a variable number of EGF-like repeats responsible for receptor–ligand interaction and (b) the negative regulatory region (NRR) with the S1 and S2 cleavage sites; followed by (c) a short transmembrane domain (TM) which contains the critical S3 cleavage side, and (d), the highly relevant NOTCH intracellular domain (NICD). This basic architecture is similar in all four NOTCH receptors, although the length of these domains and, for example, the number of EGF-like domains, varies. The functionally essential NICD of all four NOTCH receptors is itself composed of several distinguished domains, including an RBPjκ association module (RAM), ankyrin repeats (ANK; flanked by nuclear localization signal (NLS) sequences), the transactivation domain (TAD), and proline/glutamic acid/serine/threonine-rich motifs (PEST), responsible for interaction with transcription factors (RAM, ANK), regulation (TAD), and protein degradation (PEST). NOTCH ligands usually have a much simpler structure, consisting only of a single membrane-anchored extracellular domain made of EGF-like repeats region and DSL domains [1,2,3,4].
After translation, the NOTCH proteins traffic to the plasma membrane via the endoplasmic reticulum (ER) and the Golgi apparatus [5], and undergo a series of post-translational modifications (PTMs), which are essential for the specificity of ligand recognition, and subsequently, the strength and functionality of NOTCH–DSL interactions. PTMs involve, among others, the addition of O-glycans (O-glycosylation and O-xylosylation by POGLUT1, O-fucosylation by POFUT1, O-GlcNAcylation by EOGT), all linked to the EGF-like extracellular domains of the receptors [6,7]. The key factors for Notch glycosylation are the Fringe family of proteins (lunatic fringe, manic fringe, and radical fringe) [8]. These are glycosyltransferases that transfer N-acetylglucosamine to the O-linked fucose of all four Notch receptors. Fringe proteins can act both as tumor suppressors and oncogenes, depending on cancer type, cancer progression stage, and probably unknown factors such as the expression of the different fringe proteins in individual cells and tissues, or the binding of different ligands and cell–cell interactions. This illustrates the complexity of Notch signaling and its regulation. In addition, even before the insertion into the plasma membrane, the first proteolytic cleavage (S1) of the maturing NOTCH receptors occurs, via Furin-like convertase. Also, this already occurs within the compartments of the Golgi apparatus [4].
The NOTCH receptor/ligand interaction, a mechanosensory event initiated by cell-cell contacts, triggers the endocytosis of ligands, which creates a pulling force and consequently initiates the canonical Notch pathway [9,10]. In the absence of a ligand, NOTCH receptors are unable to further activate, because the activating cleavage site (S2), is in an autoinhibited conformation state inside NRR [11]. After receptor–ligand binding, the DSL-NECD complex undergoes endocytosis in the signaling sending cell. This step exerts a pulling force which allows the unfolding and exposure of the S2 site to proteolytic cleavage by ADAM protease (S2 cleavage; mainly ADAM10 and 17), followed by a “final cut” mediated by the γ secretase complex (γS; S3 cleavage). This last proteolytic cleavage results in the release of the NICD into the cytoplasm, followed by its translocation to the nucleus [2]. Once in the nucleus, the NICD binds to the DNA-binding protein RBP-J, (recombination signal binding protein for immunoglobulin kappa J region; also known as CSL, CBF-1, Su(H) or Lag-1). In the absence of the NICD, RBP-J binds several transcriptional co-repressors and epigenetic players, such as histone deacetylases (HDACs), which prevents the active transcription of Notch target genes. The presence of the NICD protein enables the dissociation of these co-repressors from RBP-J, which allows the binding of an essential co-activator, Mastermind-like protein 1 (MAML1), and the formation of a ternary complex (known as the Notch ternary complex; NTC). The NTC acts as a scaffold for binding additional transcriptional co-activators and enzymes related to DNA binding and epigenetic loosening, such as the p300 family of histone acetyltransferases (HATs, including EP300).
The high flexibility of this system, related to the possibility of dynamic interactions of the NTC with a large number of additional regulatory proteins, enables the strict control of the transcription of “classic” NOTCH target genes, such as the Hes/Hey family of transcriptional repressors, CCND1 (Cyclin D1) and MYC, by binding to their promoter or enhancer region [10,12,13,14]. It is also important to consider that the four NOTCH receptors have partially overlapping but also unique functions, and that they activate a different spectrum of target genes. The four NOTCH receptors further show markedly different cell- and tissue-type specific expression, and the spectrum of Notch target genes strongly depends on the cell and tissue type under investigation [15]. The transcriptional targets of Notch signaling are largely context-specific, and many of these genes are even located in tissue-specific enhancers [16], which allows a differential function that depends on context and even time. Thus, Notch signaling depends to a large degree on context [17] and (micro)environment, cell–cell contacts, and cell density in various tissues. As a mechanoreceptor, it is also likely that even aspects such as the composition, density, stiffness, and pressure of the extracellular matrix strongly affect Notch pathway activity, quality, strength, and the direction of signaling.
In addition, non-canonical pathways [18,19] of Notch signaling that bypass RBP-Jκ and MAML have been reported mainly in immune cells, and less frequently in solid cancers, although their mechanisms and relative importance remain debated [20,21]. Also, the non-canonical Notch pathway is associated with NICD activity, although it does not use the formation of the NTC complex. In variations of this non-canonical Notch pathway, the interaction of the NICD may occur with various signal molecules, such as AKT, mTOR, NF-κB [22], or HIF-1α [23], ultimately leading to changes in the expression of the usual broad spectrum of target genes. Interestingly, the activation of the non-canonical pathway may also be triggered by a non-canonical ligand, or it can occur in the absence of a ligand, and it may not even require NOTCH receptor cleavage [9,24]. The role of the non-canonical Notch pathway and its impact on the development and progression of cancer is still an intensively studied area that may shed new light on the possibilities of functional Notch modulation in the treatment of cancer [25] in the future. Understanding Notch signaling allows researchers to interfere with its key structures and functional connections, which may achieve specific biological effects and processes that are involved in the progression of diseases such as cancer.
Generally, Notch can play an essential, but pleiotropic, role during various stages of carcinogenesis and cancer progression, with highly tissue-specific and context-dependent outcomes, especially when considering the cross-talk with other signaling pathways, such as Wnt and Hedgehog [26]. Nevertheless, the dysregulation of Notch signaling is observed in a wide spectrum of cancers, which proved a significant incentive for interfering with it as a therapeutic target.

2. Two Sides of the Same Coin—Notch Gain- vs. Loss-of-Function

Notch signaling is an evolutionarily conserved pathway that relies on signal transduction between neighboring cells. The Notch system enables the control of the spatial localization of cells in context, and their development into organized tissue structures. During embryogenesis, Notch (together with other pathways such as Wnt and Hedgehog) is responsible for the control of somitogenesis and, furthermore, for the development of the dermis, muscles, vascular, skeleton, and nervous systems [27,28]. Notch signaling also plays a key role in maintaining tissue homeostasis after birth and maturity, by controlling processes such as tissue turnover and regeneration, wound healing, and immune response [27,29,30].
The dysregulation of Notch signaling plays an important role in cancer development. Notch signaling controls basic processes related to cancer initiation and development, such as chemo- [31,32], immune- [33,34], and radio-resistance [35], tumor cell plasticity, epithelial-to-mesenchymal transition (EMT) [36,37,38], angiogenesis [39], and the maintenance of a cancer stem-cell-like phenotype [40,41,42]. Hyperactivated Notch signaling may, thus, also play a role in tumor dormancy, relapse, and local and distant metastasis. In general, two fundamentally opposing patterns of dysregulation of Notch signaling are observed in cancers: oncogenic hyper-activation (or gain-of-function, GoF mutations), versus tumor suppressor function (due to loss-of-function, or LoF mutations). Activating and oncogenic GoF mutations in the NOTCH receptors (mostly NOTCH1) typically play a role in promoting later-stage tumor progression but can also act as gatekeepers or lineage-specific driver mutations involved in cancer initiation, such as in T-ALL [43]. Genome-wide mutation analyses from the past two decades suggest a strong, causal connection between a hyperactivated Notch pathway and the development of certain types of cancer, including acute myeloid leukemia (AML) [44], bladder cancer (BC) [45], adenoid cystic carcinoma (ACC) [46], and, above all, T-cell acute lymphoblastic leukemia (T-ALL) [43,47,48,49]. Activating NOTCH receptor mutations are also frequently found in advanced, triple-negative breast cancer (TNBC) [50,51,52,53]. Alterations in the amplification and overexpression of Notch receptors are prevalent in breast cancer. Predominantly, these alterations involve NOTCH2 (up to 13%) [54] and NOTCH1 (2–5%), while occurrences are slightly less frequent for NOTCH3 and NOTCH4 (2–3%) [55,56]. The hyperactivation and/or overexpression of NOTCH receptors typically results in the strong, ligand-independent activation of the receptor, while GoF mutations typically increase the lifespan of the functional domain of the receptor (Notch intracellular domain; NICD [3]). In addition, mutations in other genes associated with the Notch pathway, like AJUBA [57] and FBXW7 [58], can lead to the altered lifespan of NOTCH proteins and modulation of the functions of the NICD. AJUBA, for example, modulates Notch signaling by the sequestration of the NUMB protein, which promotes effective nuclear translocation of the NICD to regulate gene expression. It is related to the fact that NUMB is able to directly inhibit Notch functions by promoting the degradation and polyubiquitination of Notch proteins [59]. Surprisingly, it has recently been discovered that NUMB can enhance Notch signaling by regulating the proteasome pathway. This is related to the fact that NUMB, by controlling the activity of BAP1 deubiquitinase (BRCA1-related protein 1) and BRCA1 (Breast Cancer type 1 susceptibility protein), facilitates the stabilization of NICD [60].
FBXW7 is a F-box-containing protein that acts mainly as a tumor suppressor. FBXW7 regulates the activity of a range of substrates, including NOTCH1, through degradation by the ubiquitin-proteasome system. Functional, active FBXW7 regulates NOTCH protein activity by controlling its half-life, maintaining optimum protein levels in cells. LoF mutations of FBXW7, which are common in head and neck cancers, thus lead to an extended half-life and hyperactivity of the NICD [61]. Despite these important roles in Notch signaling, no small-molecule inhibitors have yet been developed for targeting FBXW7 or AJUBA.
LoF mutations in NOTCH receptors are characteristic of squamous cell carcinomas (SCC) such as head and neck squamous cell carcinoma (HNSCC) [62], lung squamous cell carcinoma (LSCC), and cutaneous squamous cell carcinoma (CSCC) [63,64]. They may act as gatekeeping mutations early in tumor initiation and the first stages of progression [65]. In these cases, NOTCH receptors primarily act as tumor suppressor [66] genes, as reviewed in references [17,67]. It is assumed that NOTCH receptors essentially act as tumor suppressors in squamous epithelial cells and may be required for the terminal (squamous) differentiation of the skin, cervical or oral mucosa [68]. The loss of Notch signaling likely promotes the further development of tumors in the early stages of dysplastic tissues (e.g., in virus-infected cells or cells exposed to UV light or chemical carcinogens) by impairing the balance between proliferation, terminal squamous cell differentiation, and pro-apoptotic signaling. This may, consequently, favor the maintenance and proliferation of mutated stem/progenitor cells which, along with other mutations, could then further contribute to initiating the process of carcinogenesis [69,70]. This hypothesis is supported by the observation that γ-secretase inhibitors, a class of drugs that reduce Notch signaling, may increase the risk of skin cancer development in patients who have received any of these drugs in clinical studies [71]. In HNSCC, truncating, missense, and splice site mutations in NOTCH1 that result in non-functional proteins are found in up to 20% of the cases [54]. The NOTCH1 mutation frequency is significantly higher in cancers that are negative for human papillomaviruses (HPV), indicating a different etiology of these distinct cancer subtypes [72,73]. LoF mutations in other NOTCH receptors are somewhat less frequently found, e.g., for NOTCH2 in 5–8% of HNSCC, for NOTCH3 in 3–5% of the cases, and only 2–3% in NOTCH4. Taken together, around 35–40% of all cases show aberrations in NOTCH receptors [74,75,76] (according to cBioPortal for Cancer Genomics (https://www.cbioportal.org (accessed on 10 May 2023) and NCI Genomic Data Commons (https://gdc.cancer.gov (accessed on 10 May 2023). The loss of Notch signaling in HNSCC is also reviewed in several publications [66,68,77], which also mention the potential bimodal characteristics of Notch mutations at later stages of cancer progression. The presence and role of gain-of-function alterations, such as overexpression in NOTCH receptors in HNSCC and other squamous carcinomas, remains debated. GoF mutations, such as those typical in T-ALL, are not found in HNSCC, and functional alterations are mainly related to high expression levels. Recent publications have reported a diagnostic and prognostic association of high-level NOTCH receptor expression in HNSCC [78] in European patients. Other reports of GoF alteration in Notch signaling components, including the ligands, have been predominantly observed in Asian patient cohorts, in which many suffer from different risk factors compared to European and American patients [79]. There are also reports of GoF mutations in HNSCC or other squamous carcinomas [79] in advanced, recurrent, or metastatic cancers. This suggests a potential biphasic or bimodal role of Notch pathway activity and mutations in cancer progression [77,80], in which the initially silenced pathway may be re-activated and even hyperactivated at later progression stages [80,81,82]. The relevance of this for cancer therapy and patient outcome is still poorly understood. Future research may need to also consider the aberrant expression of Notch receptors and ligands in the tumor microenvironment, such as in cancer-associated fibroblasts (CAFs), lymph vessels, and blood vessel endothelial cells. Generally, Notch signaling in different cell types within the tumor tissues is likely to significantly contribute to the intra-tumor heterogeneity, but is poorly investigated.
If additional, Notch-pathway-related genes such as AJUBA, EP300, and FBXW7 are included (each mutated with a frequency of 7–8% in HNSCC) and it appears that about two-thirds or >60% of all HNSCC harbor changes in Notch signaling [83]. Furthermore, a substantial number of tumors exhibit epigenetic silencing of the second allele of NOTCH1 located on chromosome 1 [52]. Consequently, this leads to a significant reduction in both the expression and functionality of this receptor. Epigenetic silencing is not routinely observed for the other NOTCH receptors. The degree to which the four NOTCH receptors may be able to functionally compensate for each other remains unclear, but is to be considered a possibility that requires further research. In the past, this has been subject mainly to studies in developmental biology [84], but not in oncology. In particular, NOTCH1 and 2 have been demonstrated to, at least partially, compensate for each other [85], for example, in maintaining the functional stem cell niche in gastrointestinal crypt structures [86]. There are no comparable studies in cancers.
Mutations in Notch ligands are found more rarely in the same cancers, ranging only between 0.5 and 3% for DLL1, DLL3, DLL4, JAG1, or JAG2. Most of these genetic aberrations are amplifications that can lead to the overexpression of the ligands; while inactivating or truncating mutations are very rare (<0.5%; according to The Cancer Genome Atlas Program (TCGA)—NCI; https://www.cancer.gov (accessed on 26 May 2023). In addition, about 30% of HNSCC or skin cancer cases show overexpression of JAG1 or JAG2 [87], which may be in line with the activation of the Notch pathway at least in some cases [88,89], especially aggressive, late-stage cancers, and is associated with poor patient outcome [88,89].
Finally, there is a broad group of cancers such as the brain [90], urothelial [3], or small-cell lung cancers (SCLC) [91,92] in which alterations of the Notch pathway may be functionally involved, but their precise role is ambiguous and still far from being understood. In SCLC, it is even possible that both gain- and loss-of-function mutations occur simultaneously, side by side, within different cells of the same tumor tissue. Both neuroendocrine cancer cells, and cells with more epithelial differentiation, which coexist in this cancer side by side, showed inactivating and activating NOTCH1 mutations, respectively. This resulted in the mutual support and cooperation between these cell types in tumor progression and led to markedly differential sensitivity to chemotherapy. This demonstrates how critically Notch signaling and pathway activity depends on context [10], environment, extracellular matrix, and cell–cell contacts.
Furthermore, targeting the Notch signaling pathway may likely only benefit a fraction of cancer patients with a certain tumor type. This almost entirely depends on the presence or absence and functional impairment of Notch-related mutations, the genetic background of the patients, but also context- and tissue-specific parameters within individual tumors, such as activity of stromal components. Thus, targeting Notch signaling using chemotherapy in patients represents an issue with an outstanding individualized or personalized character. Interfering with Notch signaling on behalf of cancer chemotherapy may be a prototype case for future personalized cancer medicine; each patient will likely show a unique response pattern to Notch inhibitors, especially if combined with additional targeted drugs such as PI3Kinase, AKT, mTOR, and other inhibitors [25,93]. Whether any treatment with Notch-related drugs will be beneficial for patients remains to be tested ex vivo/in vitro, for example, based on physiologically relevant tumor models, such as patient-derived organoids, explants, or xenografts. In the following paragraphs, we address the most frequently used mechanisms of action exploited by small-molecule inhibitors to interfere with Notch signaling activity.

3. Compounds Interfering with Endoplasmic Reticulum-to-Golgi and Membrane Trafficking of NOTCH Receptors

Disturbing the trafficking of premature and proteolytically processed forms of the NOTCH receptors prevents the stepwise maturation and transport of NOTCH receptors to the plasma membrane, impairing Notch signaling (Figure 2). Several small molecular weight compounds have shown significant activity in this area, particularly in inhibiting Notch signaling in cases of T-ALL models. NVS-ZP7-4 is an inhibitor of the zinc-ion transporter SLC39A7 (ZIP7) that blocks the release of Zn2+ into the cytosol from ER/Golgi structures. The treatment of T-ALL cells with NVS-ZP7-4 blocked the intracellular transport of premature NOTCH proteins via ER stress, ultimately leading to apoptosis [94]. A similar model of action related to interfering with calcium efflux has recently been described. Inhibition of the calcium sarco/endoplasmic reticulum calcium ATPase (SERCA) transporter by small-molecular weight molecules, such as Thapsigargin [95,96], and Casearin J [97] impaired the correct folding and maturation of premature NOTCH receptors, and induced G0/G1 arrest in T-ALL cells. SERCA was first identified by genomic screens as a potential target in cancer cells that harbor mutated NOTCH1 [95]. This discovery has prompted the development of novel pharmacological inhibitors targeting SERCA. These inhibitors exhibit improved drug-like characteristics, fewer off-target effects, and reduced cytotoxicity in comparison to Thapsigargin. The compound CAD204520 is a small-molecule inhibitor of SERCA with reduced off-target Ca2+ toxicity that is currently in clinical trials against T-ALL [98]. The most recent developments of SERCA inhibitors targeting Notch signaling are summarized by Pagliaro et al., 2021 [99] and 2020 [100]. Other calcium transporter blockers, such as the non-selective drug Bepridil, are also suspected to show similar features [101] as described for Thapsigargin.
The inhibition of NOTCH secretion before its exit from the ER and entry into the Golgi apparatus was observed with dihydropyridine derivatives such as FLI-06 [102]. FLI-06 effectively inhibits Notch signaling in lung adenocarcinoma [103] and HNSCC [104,105]. FLI-06 treatment is associated with G1 cell cycle arrest and programmed cell death/apoptosis. The inhibition of NOTCH receptor trafficking has also been reported for Chloroquine, originally developed as one of the first antimalarial drugs, showing the ability to inhibit lysosome function and autophagy. Along these lines, the chloroquine-induced accumulation of the NOTCH1 receptor in the autophagosome was observed [106]. Similarly, Bafilomycin A1 (BafA1), a highly specific Vacuolar H+ ATPase (V-ATPase) inhibitor, leads to the accumulation of NOTCH in the endo-lysosomal system, and consequently, reduces Notch signaling in normal breast cells. Interestingly, in a breast cancer model, BafA1 treatment reduces the growth in cells expressing membrane-tethered, constitutively active NOTCH1 forms, while sparing cells expressing cytoplasmic forms [107].
Blocking membrane trafficking strongly inhibits Notch signaling due to the reduced exposure and density of NOTCH receptors on the cell surface. However, these effects are mostly related to the indirect reduction in all four types of NOTCH receptors. Not only does pan-NOTCH inhibition (blocking all four types of NOTCH receptors) complicate the research on the role of individual types of receptors, it also makes it more difficult to achieve the desired therapeutic effects. Additionally, membrane trafficking inhibition, the blocking of ion channels, and interfering with intracellular ion concentrations are all notoriously non-specific mechanisms of action. This line of research carries a high risk of non-specific interference with the transport of many other membrane proteins, and ion-dependent signaling pathways, thus making the interpretation of the results related to Notch signaling difficult.

4. NOTCH–Ligand Interaction, Γ-Secretase, and ADAM Proteases Inhibition

Notch signaling can also be modulated by modifying the sugar residues of EGF-like extracellular domains of the NECD. The inhibition of Notch signaling can be accomplished with O-fucose analogs such as 6-alkynyl and 6-alkenyl fucose, specifically by incorporating these analogs into EGF-like domain 8 (EGF8), which disrupts the DLL-NECD interaction. Interestingly, both of these analogs do not disturb Jagged-induced, Notch signaling [108]. This observation opens up a new field of study for different O-glycans to fine-tune the strength of Notch signaling by the modification of the sugar residues of the NICD.
The most widely used mechanism of action that is employed by small-molecule inhibitors relates to the proteolytic processing of NOTCH receptors (Figure 3), especially by the gamma secretase complex. γ-Secretase (γS) is a multifactor complex formed by presenilin 1/2 (homolog PS1 or PS2), nicastrin (NCT), anterior pharynx-defective 1A/1B (APH-1a or APH-1b), and presenilin enhancer 2 (PEN-2) proteins. Anchored in the cell membrane, γS is responsible for the hydrolysis of substrates through the activity of its catalytic subunit, presenilin. The proteolytic cut of the S3 cleavage site of all four NOTCH receptors represents a key event in the activation of canonical Notch signaling [109,110]. Thus, the inhibition of the γ-secretase complex by γ-secretase inhibitors (γSI or GSI) is the most intensely studied and best understood mechanism, and also the most frequently used method to attenuate Notch signaling in pharmacological research and therapy. This is true despite the observation that the γ-Secretase complex targets approximately 140 different substrates [111] and is involved in the processing of many other proteins beside NOTCH receptors, although this may be the primary function with which this class of compounds is now associated. The initial target for the first GSI, however, was Alzheimer’s disease (AD), in which the γ-secretase complex is associated with the deposition of senile plaques, resulting from the aberrant deposition of amyloid precursor protein (APP) and the production of short amyloid-β peptides (Aβ) [112]. However, GSI have not proven beneficial for the treatment or prophylaxis against AD, and all clinical trials have failed. The first GSI to be used in clinical trials against cancer in the year 2006 was, therefore, a drug repurposed from AD research, MK-0752 [113]. The drug was only partially successful, and the clinical trial overall was a disappointment; MK-0752 was a non-specific, broad-band GSI and showed only limited efficacy in inhibiting T-ALL proliferation. Nevertheless, it resulted in limited disease regression in a small number of patients. Since 2006, many more clinical trials have been conducted with a spectrum of different GSI, most of which have failed, mainly due to high cytotoxicity and/or poor efficacy. Many initial clinical phase 1 studies, aiming to find the optimal and safe drug dosage for patients, have shown excessive toxicity in patients treated with GSI, mainly due to severe gastrointestinal adverse events, often resulting in the discontinuation of these trials. Many patients enrolled in these clinical trials suffered from persistent, severe diarrhea in a dose-dependent manner [114]. The basis for this toxicity remains incompletely understood. It is likely related to the existence of multiple known and validated γ-secretase substrates in addition to APP or NOTCH1, such as ephrinB2, ERBBP4, E-cadherin, and CD44 [115], which are widely expressed in human tissues, especially the gastrointestinal tract, and some of which may also be involved in the regulation of cell renewal or carcinogenesis [116,117]. As a result, the direct on-target inhibition of Notch signaling by GSI can result in severe gastrointestinal side effects, which can be persevered by additional, off-target effects. Notch signaling regulates the differentiation of progenitor cells and/or the maintenance of stem cell status in the human intestinal epithelium. Consequently, the blocking of Notch signaling by GSI can induce clinical, endoscopic, and histological abnormalities in the gastrointestinal tract, including secretory cell metaplasia of the intestinal epithelium which, in turn, leads to diarrhea [118]. Additionally, the inhibition of Notch signaling by GSI results in the loss of crypt base columnar (CBC) stem cells and an increase in secretory goblet cells’ population at the expense of proliferating cells. Consequently, γ-Secretase can result in altered homeostasis in the gastrointestinal system due to exposure to GSI, leading to rapid weight loss and even death, consistent with a failure of tissue replenishment and lack of nutrient absorption [119]. This has significantly limited the use of GSI for the treatment of cancer patients, forcing clinicians to utilize low, suboptimal, and intermittent dosing, thus decreasing the potential drug efficacy. Some of the toxicity observed in early clinical trials can also be alleviated by co-treatment with glucocorticoids. The problem was intrinsic to first-generation GSI, including drugs like AL101 and its follow-up development AL102 [120,121], a potent, orally administered γ-secretase inhibitor, but also Crenigacestat, and Nirogacestat. All of these drugs showed similar gastrointestinal toxicities, often coupled with anemia, fatigue, or thrombocytopenia [122,123]. Even today, there is still no strong evidence for their benefit in clinical use in anticancer therapy, which has been demonstrated in a series of recent clinical trials with Crenigacestat [123,124,125,126] and other GSI [127] (see Table 1 for an overview of the recent and ongoing clinical trials). This is combined with a frequent lack of adequate postoperative response. In addition, it has been shown that the actual effectiveness of some next-generation GSI, such as Semagacestat, in specifically blocking the function of the intended target γS may be much lower than expected [128]. However, there is still hope, and trials with GSI continue to this day, as illustrated in Table 1. Recently, Nirogacestat has shown a slightly better therapeutic outcome and better tolerance, and resulted in a significant increase in progression-free survival rate and lower side effects in patients with desmoid tumors [122].
Despite the disappointment of clinical trials, and due to the continuous high demand for useful tools in academic and preclinical research, a considerable number of small molecular weight compounds have been developed since, approximately, the year 2000, all of which inhibit γ-secretase and are now successfully used in laboratory practice. The GSI now represents a large and quite heterogeneous group of small molecular weight compounds, including peptide analogs and non-peptide small-molecule drugs, that share the capacity to inhibit the activity of the γ-secretase complex [129]. According to the National Library of Medicine database (National Institutes of Health; https://www.ncbi.nlm.nih.gov (accessed on 26 May 2023), there are currently 38 registered compounds described as γ-secretase modulators. The mechanism of GSI action is mostly related to the competitive binding and blocking of the active site of presenilin for its substrates. Alternatively, compounds may interfere in a non-competitive fashion with the structure and formation of the γ-secretase complex. GSI are commonly used as pan-Notch inhibitors in basic research, e.g., to investigate the role of Notch signaling in cancer development and progression. It is worth remembering that GSI also block all other, non-cancer-related, Notch-related processes and disorders. They are now commonly used as pan-Notch inhibitors in fundamental research related to the impact of Notch signaling in tissue development, cell fate decisions, differentiation, cell signaling, but also in stress response and various disorders [130,131,132,133]. One of the most commonly used GSI is the dipeptide DAPT, a presenilin inhibitor, which acts by binding to the catalytic pocket of PS, consequently blocking the internalization of the substrate and PS activation [134]. DAPT has been successfully used as a model compound (the “gold standard”) in many studies in which the inhibition of Notch signaling is suspected to have an anti-cancer effect. The use of DAPT includes studies on osteosarcoma [135], breast [136,137,138,139], ovarian [140], cervical [141,142], prostate [143], gastric [144,145], colorectal [146], and tongue [147] cancer, making it one of the most popular GSI used in cancer research.
In addition to DAPT, there are several extensively studied and tested GSI used to prevent NOTCH activation in cancer, including benzazepines such as Dibenzazepine (DBZ; YO-01027) [148,149,150], Crenigacestat (LY3039478) [151,152], LY411575 [153], and RO4929097 [154,155], sulfonamide derivatives such as Avagacestat (BMS-708163) [156,157], MK-0752 [158] and Venetoclax [159], or even natural compounds such as Hesperidin, a polyphenolic glycoside flavonoid that has shown anticancer potential in a colon cancer model [160]. In a broader sense, this further includes additional natural compounds such as Curcumin [161] or Evodiamine, an indole alkaloid that inhibits NOTCH3 signaling in a lung cancer model [162]. In most experimental studies, these compounds are used similarly to DAPT to inhibit γ-secretase activity and, as a consequence, Notch signaling [163,164]. However, some of the newer compounds may act with greater affinity and efficacy for γS, and can therefore be used at a significantly lower concentration, for example, as reported for LY411575 [165]. Additionally, the concomitant inhibition of Notch and tyrosine kinase receptor DDR1 (Discoidin Domain Receptor 1) signaling by LY411575 and 7rh, respectively, was demonstrated to trigger the blockade of cell growth and to induce apoptosis in KrasG12V-driven tumors via the modulation of MAPK signaling. Additionally, the combination of LY411575 and 7rh induced higher apoptosis than either agent alone [166]. This shows the potential of combining Notch inhibitors with other compounds to achieve the synthetic lethality phenomenon.
Several studies suggest the usefulness of GSI in anticancer treatment based on in vivo and in vitro cancer models. For example, Crenigacestat has demonstrated anticancer activity in a model of intrahepatic cholangiocarcinoma by inhibiting tumor cell growth, oncogenic activity of CAFs, and blood vessel formation [152,167]. Interestingly, in some cases, the use of GSI is associated with a simultaneous reduction in the cytotoxicity of other drugs and the promotion of cellular regeneration processes. The inhibition of γS by Dibenzazepine has been demonstrated to be protective against cisplatin-induced nephrotoxicity [168], spleen injuries [169], paracetamol-induced hepatotoxicity [170], and neomycin-induced cochleae [171] damage. This makes Dibenzazepine an interesting compound for research in adverse events and off-target effects, e.g., by reducing the non-specific cytotoxicity of anticancer drugs like the widely used cisplatin, along with the attenuation of Notch signaling.
Unfortunately, many GSIs are still considered to be of low efficacy and low specificity as anti-cancer drugs, coupled with a high risk of poor tolerance and severe side effects. However, their wide spectrum of activity against multiple targets seems, simultaneously, to be one of their greatest advantages and disadvantages in therapy. A reasonable and ethical solution might be to use GSI in a pre-selected cohort of patients with known Notch mutation status, which would only allow the inclusion of patients with confirmed hyperactive, oncogenic Notch signaling. However, this may be difficult, in practice; there are no accepted biomarkers that would indicate overall Notch pathway activity in cells and tissues and this is further complicated by the possibility that Notch pathway activity may be significantly affected by other genes than the Notch receptors themselves.
Another solution may be the use of natural organic compounds, usually characterized by lower toxicity and side effects. Curcumin is a natural polyphenol with pleiotropic effects on the functions and activity of transcription factors (STAT, NF-κB PPAR-γ, β-catechin), growth factors (VEGF, TGF-β1), protein kinases (MAPK, EGFR, JNK, IKK), and other proteins (Bcl-2, cMyc, cyclin D1) [172,173]. Additionally, the ability of Curcumin to inhibit γS has been demonstrated [161]. Despite data that show that Curcumin may act as a Notch signaling activator [174], or that it can reverse the effects of Notch inhibition [175] in normal (non-transformed) cells, research suggests its potent inhibitory effect on Notch signaling in cancer cells. The exposure of cells and tissues to Curcumin decreases Notch activation in T-ALL [176], thymic carcinoma [177], lung cancer [178], prostate cancer [179], or melanoma [180] cells.
Compounds like Curcumin are also known as “frequent hitters” that pop up in many pharmacological screens and have, therefore, been classified as “Pan-Assay INterference compoundS (PAINS)”. Other compounds that are candidates for PAINS are EGCG (epigallocatechin-3-gallate), Sulforaphane, Resveratrol, and Genistein. These are sometimes called “the big five” phytochemicals. All these compounds show remarkable polypharmacology and may bind to or interfere with many different substrates besides Notch, thus posing a high risk of causing false-positive assay signals. Nevertheless, they have in common that all of them have been associated with targeting cancer stem cells and EMT [181,182]. In addition to the Wnt/β-catenin, Hedgehog, PI3K/Akt/mTOR, and STAT3 pathways, the Notch pathway is another crucial molecular pathway that is prominently active in cancer stem cells [183]. The same pathways have been associated with EMT, relapse, dormancy, and increased tumor cell plasticity. Despite understandable doubt or outright criticism based on their poor target specificity, several recent clinical trials have demonstrated the therapeutic efficacy of these five phytochemicals. They have also been successfully combined with other cancer therapeutics [184] and in various types of cancer. Even if Notch signaling may be only one of several candidate pathways targeted by these “big five” phytochemicals, it may yet be worthwhile to seriously consider their capacity to target aggressive cancer stem cells, which are not affected by most chemotherapeutic drugs. Furthermore, it has been shown that derivatives of curcumin, so-called curcuminoids, can show much more specific effects on Notch pathway inhibition [176], suggesting that the pharmacophore of curcuminoids may, indeed, be promising for future medicinal chemistry and the development of more specific agents.
Less frequent are the pharmacological attempts to inhibit Notch signaling by blocking the S2 cleavage step, performed by ADAM proteases (ADAM-10, ADAM-17). For the same reasons as outlined for GSI, these compounds are also characterized by pleiotropic effects on multiple targets. They are not restricted to Notch signaling, but also inhibit, for example, matrix metalloproteinases like MMP2 and MMP9 [185]. The effective inhibition of ADAM-17, resulting in a reduction in Notch signaling by Marimastat (carboxamide), was observed in renal carcinoma [186] and by the compound ZLDI-8 in non-small-cell lung cancer [187] and colorectal cancer [188]. In addition, INCB7839 (Aderbasib), a dual-specific inhibitor targeting both ADAM10 and 17, has been tested in clinical trials against pediatric glioma. It appears, however, that this compound generally inhibits the shedding of many ligands required for cell–cell signaling, and has also been active, for example, against EGF receptor signaling [189,190]. In some cases, the results of inhibition of ADAM-17 seem to reflect the results of inhibition of GSI, i.e., inhibition of tumor cell growth and mobility as well as induction of apoptosis. However, the relatively small amount of reliability in this area makes it difficult to assess the usefulness of inhibiting ADAM proteases in the inhibition of Notch signaling and anticancer therapy.

5. NICD-Dependent Transcription in the Nucleus

The assembly and operation of the NICD/RPB-J/MAML transcription complex (a.k.a. NTC) is essential for the transcriptional initiation of Notch target genes (Figure 4). Several small molecular weight compounds have been developed that specifically modify the functionality of the complex. They act mainly by preventing individual subunits from binding to the complex. In most cases, this leads to the inhibition of the expression of genes dependent on Notch signaling. For example, the rhodanine ester compound IMR-1 (Mastermind Recruitment Inhibitor-1), and its metabolite IMR-1A, were demonstrated to interfere with the recruitment of MAML1 to RPB-J [191], consequently blocking the formation of the NTC and downstream gene expression in an esophageal adenocarcinoma model. Similarly, an α-helical peptide derived from mastermind-like 1 (SAHM1) acts as a dominant-negative MAML1 mimetic. Like IMR-1 and IMR-1A, SAHM1 attenuates Notch signaling by blocking the MAML/RPB-J interaction and its antitumor effects have been reported against bladder cancer [192] and acute myeloid leukemia [193]. However, SAHM1 showed no significant effects on B-cell acute lymphoblastic leukemia (B-ALL) [194]. The synthetic molecule CB-103 blocks the functional interaction between RBP-J and NICD, showing antitumor activity against T-ALL cells, TNBC xenografts in mouse models [195], and T-ALL in clinical studies [196]. Additionally, CB-103 has shown the ability to inhibit Notch signaling in other cancer models, including clear-cell renal cell carcinoma [197]. Presently, CB-103 is in at least one Phase 2 clinical study (Study of CB-103 in Adult Patients with Advanced or Metastatic Solid Tumours and Haematological Malignancies; ClinicalTrials.gov Identifier: NCT03422679, see Table 1).
The downregulation of NOTCH receptors has also been shown with pan-HDAC inhibitors such as Givinostat and Trichostatin A in T-ALL. More detailed analysis showed that this effect is related to the inhibition of HDAC6, which can also be achieved by using more specific inhibitors, such as Tubacin [198]. HDACs, along with RBP-J and other co-repressor proteins, are components of the inactive transcription complex, which blocks the expression of Notch-targeted genes. Additionally, the inhibition of HDAC6 has been shown to lead to the disruption of NOTCH3 receptor transport. This leads to the increased accumulation and degradation of NOTCH3 in lysosomes, which results in the attenuation of Notch signaling [198]. This shows a direct relationship between the functions of the NICD-dependent nuclear transcription complex and membrane trafficking of NOTCH receptors.

6. Small Molecular Weight Compounds with Pleiotropic Functions in Relation to Notch Signaling

The activation of Notch signaling has shown some beneficial outcomes in studies of certain types of cancer. The possibility of the activation of genes dependent on Notch signaling, because of the modulation of the previously described Notch ternary complex, was confirmed. Very often, however, the compounds used for this purpose show a context-dependent effect.
RIN-1 (RBPJ Inhibitor-1; 2-(2-Fluorophenoxy)-4-(1-methyl-1H-pyrazol-5-yl) benzamide) is an inhibitor that blocks the functional interaction of RBP-J with SHARP (also known as SMRT/HDAC1-associated repressor protein or MINT-Msx2-interacting nuclear target) [199]. SHARP is a scaffold protein that functions in multiple ways: (1) SHARP acts as an NTC co-repressor by binding NCoR1/2 (Nuclear receptor co-repressor 1 and 2) and HDAC, consequently inhibiting transcription. In addition, (2) SHARP is also an NTC co-activator that acts by recruiting KMT2D lysine methyltransferases [200,201] to the complex. Despite the bimodal function of SHARP, studies suggest that blocking SHARP/RPB-J interaction by RIN-1 leads to an increase in the expression of genes dependent on Notch signaling [199]. Therefore, RIN-1 can be generally considered an activator of Notch signaling.
Valproic acid (VPA), an HDAC inhibitor, is a recognized stimulator of Notch signaling and exhibits anticancer properties in various cancer types, such as cervical cancer (resulting in cell cycle arrest, proliferation suppression, and apoptosis induction) [202] and human pancreatic carcinoid (as above) [203]. On the other hand, it has been shown that VPA can also attenuate Notch signaling, as was the case with chronic myelogenous leukemia K562 cells [204], and hepatocellular carcinoma HepG2 [205] cells. In addition to HDAC inhibition (selectivity to HDAC6 and HDAC8), VPA works by enhancing γ-aminobutyric acid (GABA) signaling and inhibiting calcium and sodium channels [206], making it difficult to use as a selective inhibitor of Notch signaling. Moreover, a significant limitation of the use of VPA may be its relatively low efficacy: concentrations usually exceeding 4 mM are necessary to obtain a measurable modulation of Notch signaling.
Another widely used compound that can both activate or attenuate Notch signaling is Resveratrol, one of the “Big Five” natural compounds. Interestingly, antitumor activity was reported for both modes of action; Resveratrol (and other polyphenols) are well-known activators of Notch signaling in normal, non-transformed cells, including endothelial cells [207], but also some types of cancer, including osteosarcoma [208], neuroendocrine carcinoid [209], and thyroid cancer [210,211]. However, the currently available data on the putative activation of NOTCH signaling in cancer by Resveratrol are ambiguous and suggest that the compound may suppress Notch signaling. For example, an anticancer effect accompanied by the downregulation of proteins, such as NOTCH1, Jagged1, DLL4, and HES5, by Resveratrol was observed in MDA-MB-231 breast cancer cells [212]. In OVCAR-3 ovarian cancer cells, the downregulation of NOTCH-2 and HES1 was observed [213] while, in the cell lines A2780 and SKOV3, a reduction in full and cleaved NOTCH1 was observed [214]. In some cases, the net effect of Resveratrol does not allow the determination of its mechanism of action as an activator/inhibitor of Notch signaling, as was the case with Glioblastoma multiforme (GBM) cell lines U87MG and T98G, in which a simultaneous decrease in NOTCH1 and NICD expression and an increase in HES1 expression was observed [155]. Resveratrol is also known as a pan-inhibitor of histone deacetylases in cancer cells [215], which may explain its potential to modulate Notch signaling. In addition, Resveratrol interacts with elements of the various signaling pathways, such as β-catenin, Smad2/3, IKK, AKT, PI3K, STAT3, and HIF-1α, which are closely related to canonical and non-canonical Notch signaling modulation [216]. Consequently, the result of its action is, again, pleiotropic and often indirect, similar to that described for Curcumin, and it is difficult to determine the efficacy and specificity of these compounds in relation to Notch signaling.
Another broad cellular mechanism that has been exploited to target NOTCH1 protein levels within cancer cells is the use of proteasome inhibitors, such as Bortezomib. Bortezomib effectively represses the transcription of NOTCH1, thus blocking the activation of Notch downstream effectors including HES1, GATA3, or RUNX3, and resulting, for example, in the downregulation and dissociation of the major transcriptional activator Sp1 from the NOTCH1 promoter. This causes significant cytotoxicity in T-ALL cells [217] which is partly overcome by the overexpression of the NOTCH1 intracellular domain (NICD). Furthermore, a combination of Bortezomib with the histone deacetylase inhibitor romidepsin further increased the cytotoxic effects [218], at least in a mouse model of leukemia. Interestingly, Bortezomib has been shown to increase the expression of the Notch-dependent genes downstream in a γ-secretase-independent manner in tumor-draining CD8+ T cells [219]. This demonstrates the significant context-dependent influence of proteasome activity on Notch signaling. It has also been debated whether part of the effects of GSI in breast cancers may be related to targeting the proteasome [220], effects that have also been observed in neuroblastoma. It remains unclear if the effects of proteasome inhibitors against cancers may partly relate to the blocking of Notch signaling, and vice versa, if the anti-cancer effects of GSI may, at least partly, stem from the inhibition of the proteasome. It should also be mentioned that, Fostamatinib is functionally related to proteasome activity via NUMB activity, which blocks calcium-/calmodulin-dependent protein kinase CAMK1. CAMK1 regulates NUMB-mediated endocytosis by the phosphorylation of NUMB on the Serine residue 276 and 295; interfering with this phosphorylation effectively blocks the nuclear import of the NICD (for reference, see GenBank, https://go.drugbank.com/drugs/DB12010 (accessed on 26 May 2023).
Last but not least, an interesting concept has been introduced that specifically targets mutated tumor suppressor genes with nonsense mutations. Apart from TP53, NOTCH1 and FAT1 also show high frequencies of nonsense mutations, especially in HNSCC. The treatment of NOTCH-mutated HNSCC cell lines with the compound PTC-124 (Ataluren) was demonstrated to re-express the functional NOTCH1 protein and induce the expression levels of Notch-regulated genes such as HES1, HES5, and AJUBA, thus effectively reverting the loss-of-function mutations found in the tumor cells [221]. PTC-124 selectively induces the ribosomal readthrough of premature but not normal termination (stop) codons. Mechanistically, PTC-124 acts by binding to rRNA, which inhibits the ribosomal activity to mRNA termination. Consequently, PTC-124 mostly allows the “bypassing” of UGA nonsense mutations in the mRNA. This makes the synthesizing of a full-length protein possible with a changed amino acid residue at the mutation site [222,223,224,225]. PTC-124 was originally developed for the treatment of cystic fibrosis, where it was also able to functionally rescue some nonsense mutations in the CFTR transporter gene [226]. Its application in cancer is a classic example of drug repurposing. Additional examples with relevance for Notch signaling have been described in the recent literature [159], not only identifying the Bcl-2 inhibitor Venetoclax as a γ-secretase inhibitor, but also indicating its superior efficacy over other “standard” GSI such as RO4929097. Another example is the identification of the FDA-approved antibiotic Fidaxomicin as a potential RBP-J inhibitor [227]. Fidaxamicin was subsequently shown to block RBP-J-dependent transcription and thereby inhibit the Notch-dependent growth of TNBC. Similarly, the lipid-lowering agent Lomitapide was identified as a potent compound targeting proteolytic enzymes, most prominently ADAM17 and the γ-secretase complex [228]. It showed remarkable proliferation-blocking, antitumor properties, especially against TNBC cells and cell lines. Lomitapide also induced significant apoptosis and blocked cell cycle progression in TNBC cells. Another concept that may be relevant here is the use of synthetic lethality to target cancer cells that have mutated (loss-of-function) Notch receptors. This was suggested over 10 years ago [229], but only recently has the concept been experimentally demonstrated in oral cancers. A broader acceptance of this strategy might be very useful in developing more successful combination therapies, either of targeted compounds specifically targeting tumors with functional loss of Notch signaling, or vice versa, using Notch inhibitors to target cancer cells that have losses in other tumor suppressor genes. A thorough investigation of patient-specific pathway activity is the condition for this concept to work in personalized cancer medicine, but the initial steps have been taken and Notch inhibitors may be “on board”.
Currently, selective activators of Notch signaling are much more difficult to obtain (and validate) than inhibitors. This leads to several inconveniences when planning experiments, including the selection of appropriate cancer cell lines with clearly traceable Notch activity, as well as assessing the likely interaction of the tested compounds with other signaling pathways.
A summary of the clinical trials for small-molecule compounds targeting Notch signaling is presented in Table 1.

7. Conclusions

The development of research on small-molecule modulators of Notch signaling offers new perspectives for a better understanding of the processes behind cancer initiation and progression, including processes that lead to poor patient survival, such as relapse, acquired drug resistance, increased aggressive behavior, and metastasis. Currently, Notch small-molecule modulators are successfully used and are relatively convenient and reliable in laboratory practice, despite the notorious question related to their specificity. Nevertheless, they have advantages over recombinant monoclonal antibodies that either target Notch receptors or ligands, such as Brontictuzumab, a therapeutic antibody targeting, specifically, NOTCH1 [230,231,232]. The realistic potential of Notch pathway modulators (inhibitors and a few activators) in basic research, applied medicine, and eventually, in clinical trials is still under consideration. Despite significant recent progress in this field, especially in the identification of targets besides the γ-secretase complex and the proteolytic processing of the Notch receptors themselves, the biggest problem appears to be the low selectivity of most compounds. This can be alleviated by increasingly targeting the Notch receptor and addressing the need for the careful stratification of patient cohorts in order to identify those candidates that may target downstream signaling events, such as the NICD/RBP-J/MAML1 complex and transcriptional activity, instead of directly targeting the Notch receptors themselves. In clinical treatment, obtaining appropriate Notch modulators that are capable of interfering with Notch signaling without excessive off-target and non-specific side effects represents an urgent, unmet clinical need, but also a near-term, realistic goal for medicinal chemistry in collaboration with translational cancer research and personalized medicine.

Author Contributions

Conceptualization, A.C. and M.N.; methodology, A.C.; software, J.K.; validation, M.N. and A.C.; formal analysis, A.C., M.N. and J.K.; investigation, A.C. and M.N.; resources, M.N.; data curation, A.C. and M.N.; writing—original draft preparation, A.C.; writing—review and editing, M.N., A.C. and J.K.; visualization, J.K.; supervision, M.N.; project administration, A.C.; funding acquisition, M.N. and J.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Polish National Science Centre (NCN): UMO-2020/37/B/N24/03920 and DEC-2021/41/N/NZ5/01938.

Institutional Review Board Statement

The study was conducted according to the guidelines of and approved by the Institutional Review Board of the Medical University of Lublin.

Data Availability Statement

The data presented in this study are available in the presented article. Any additional data related to this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish.

References

  1. Sherry, K.P.; Das, R.K.; Pappu, R.V.; Barrick, D. Control of transcriptional activity by design of charge patterning in the intrinsically disordered RAM region of the Notch receptor. Proc. Natl. Acad. Sci. USA 2017, 114, E9243–E9252. [Google Scholar] [CrossRef]
  2. Steinbuck, M.P.; Winandy, S. A Review of Notch Processing with New Insights Into Ligand-Independent Notch Signaling in T-Cells. Front. Immunol. 2018, 9, 1230. [Google Scholar] [CrossRef] [PubMed]
  3. Ferreira, A.; Aster, J.C. Notch signaling in cancer: Complexity and challenges on the path to clinical translation. Semin. Cancer Biol. 2022, 85, 95–106. [Google Scholar] [CrossRef] [PubMed]
  4. Kopan, R.; Ilagan, M.X.G. The Canonical Notch Signaling Pathway: Unfolding the Activation Mechanism. Cell 2009, 137, 216–233. [Google Scholar] [CrossRef] [PubMed]
  5. Apodaca, G.; Brown, W.J. Membrane traffic research: Challenges for the next decade. Front. Cell Dev. Biol. 2014, 2, 52. [Google Scholar] [CrossRef] [PubMed]
  6. Antfolk, D.; Antila, C.; Kemppainen, K.; Landor, S.K.-J.; Sahlgren, C. Decoding the PTM-switchboard of Notch. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 118507. [Google Scholar] [CrossRef] [PubMed]
  7. Li, Z.; Fischer, M.; Satkunarajah, M.; Zhou, D.; Withers, S.G.; Rini, J.M. Structural basis of Notch O-glucosylation and O–xylosylation by mammalian protein–O-glucosyltransferase 1 (POGLUT1). Nat. Commun. 2017, 8, 185. [Google Scholar] [CrossRef]
  8. Mugisha, S.; Di, X.; Disoma, C.; Jiang, H.; Zhang, S. Fringe family genes and their modulation of Notch signaling in cancer. Biochim. Biophys. Acta Rev. Cancer 2022, 1877, 188746. [Google Scholar] [CrossRef]
  9. Majumder, S.; Crabtree, J.S.; Golde, T.E.; Minter, L.M.; Osborne, B.A.; Miele, L. Targeting Notch in oncology: The path forward. Nat. Rev. Drug Discov. 2021, 20, 125–144. [Google Scholar] [CrossRef]
  10. Bray, S.J. Notch signalling in context. Nat. Rev. Mol. Cell Biol. 2016, 17, 722–735. [Google Scholar] [CrossRef]
  11. Gordon, W.R.; Zimmerman, B.; He, L.; Miles, L.J.; Huang, J.; Tiyanont, K.; McArthur, D.G.; Aster, J.C.; Perrimon, N.; Loparo, J.J.; et al. Mechanical Allostery: Evidence for a Force Requirement in the Proteolytic Activation of Notch. Dev. Cell 2015, 33, 729–736. [Google Scholar] [CrossRef] [PubMed]
  12. 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]
  13. Nam, Y.; Sliz, P.; Pear, W.S.; Aster, J.C.; Blacklow, S.C. Cooperative assembly of higher-order Notch complexes functions as a switch to induce transcription. Proc. Natl. Acad. Sci. USA 2007, 104, 2103–2108. [Google Scholar] [CrossRef] [PubMed]
  14. Kovall, R.A.; Blacklow, S.C. Mechanistic insights into notch receptor signaling from structural and biochemical studies. Curr. Top. Dev. Biol. 2010, 92, 31–71. [Google Scholar] [PubMed]
  15. Orzechowska, M.; Anusewicz, D.; Bednarek, A.K. Functional Gene Expression Differentiation of the Notch Signaling Pathway in Female Reproductive Tract Tissues—A Comprehensive Review with Analysis. Front. Cell Dev. Biol. 2020, 8. [Google Scholar] [CrossRef]
  16. Castel, D.; Mourikis, P.; Bartels, S.J.J.; Brinkman, A.B.; Tajbakhsh, S.; Stunnenberg, H.G. Dynamic binding of RBPJ is determined by notch signaling status. Genes Dev. 2013, 27, 592616. [Google Scholar] [CrossRef]
  17. Misiorek, J.O.; Przybyszewska-Podstawka, A.; Kałafut, J.; Paziewska, B.; Rolle, K.; Rivero-Müller, A.; Nees, M. Context matters: Notch signatures and pathway in cancer progression and metastasis. Cells 2021, 10, 94. [Google Scholar] [CrossRef]
  18. Dongre, A.; Surampudi, L.; Lawlor, R.G.; Fauq, A.H.; Miele, L.; Golde, T.E.; Minter, L.M.; Osborne, B.A. Non-canonical Notch signaling drives activation and differentiation of peripheral CD4+ T cells. Front. Immunol. 2014, 5, 54. [Google Scholar] [CrossRef]
  19. Charbonnier, L.M.; Wang, S.; Georgiev, P.; Sefik, E.; Chatila, T.A. Control of peripheral tolerance by regulatory T cell-intrinsic Notch signaling. Nat. Immunol. 2015, 16, 1162–1173. [Google Scholar] [CrossRef]
  20. Ayaz, F.; Osborne, B.A. Non-canonical Notch signaling in cancer and immunity. Front. Oncol. 2014, 4, 345. [Google Scholar] [CrossRef]
  21. Jin, S.; Mutvei, A.P.; Chivukula, I.V.; Andersson, E.R.; Ramsköld, D.; Sandberg, R.; Lee, K.L.; Kronqvist, P.; Mamaeva, V.; Östling, P.; et al. Non-canonical Notch signaling activates IL-6/JAK/STAT signaling in breast tumor cells and is controlled by p53 and IKKα/IKKβ. Oncogene 2013, 32, 4892–4902. [Google Scholar] [CrossRef] [PubMed]
  22. Hossain, F.; Sorrentino, C.; Ucar, D.A.; Peng, Y.; Matossian, M.; Wyczechowska, D.; Crabtree, J.; Zabaleta, J.; Morello, S.; Del Valle, L.; et al. Notch signaling regulates mitochondrial metabolism and NF-κB activity in triple-negative breast cancer cells via IKKα-dependent non-canonical pathways. Front. Oncol. 2018, 8, 575. [Google Scholar] [CrossRef] [PubMed]
  23. Lakhan, R.; Rathinam, C.V. Deficiency of Rbpj Leads to Defective Stress-Induced Hematopoietic Stem Cell Functions and Hif Mediated Activation of Non-canonical Notch Signaling Pathways. Front. Cell Dev. Biol. 2021, 8, 622190. [Google Scholar] [CrossRef] [PubMed]
  24. Conner, S.D. Regulation of Notch Signaling Through Intracellular Transport. Int. Rev. Cell Mol. Biol. 2016, 323, 107–127. [Google Scholar] [PubMed]
  25. Johnson, F.M.; Janku, F.; Gouda, M.A.; Tran, H.T.; Kawedia, J.D.; Schmitz, D.; Streefkerk, H.; Lee, J.J.; Andersen, C.R.; Deng, D.; et al. Inhibition of the Phosphatidylinositol-3 Kinase Pathway Using Bimiralisib in Loss-of-Function NOTCH1-Mutant Head and Neck Cancer. Oncologist 2022, 27, 1004-e926. [Google Scholar] [CrossRef]
  26. Kumar, V.; Vashishta, M.; Kong, L.; Wu, X.; Lu, J.J.; Guha, C.; Dwarakanath, B.S. The Role of Notch, Hedgehog, and Wnt Signaling Pathways in the Resistance of Tumors to Anticancer Therapies. Front. Cell Dev. Biol. 2021, 9, 857. [Google Scholar] [CrossRef]
  27. Siebel, C.; Lendahl, U. Notch signaling in development, tissue homeostasis, and disease. Physiol. Rev. 2017, 97, 1235–1294. [Google Scholar] [CrossRef]
  28. Reichrath, J.; Reichrath, S. Notch Signaling and Embryonic Development: An Ancient Friend, Revisited. Adv. Exp. Med. Biol. 2020, 1218, 9–37. [Google Scholar]
  29. Kiyokawa, H.; Morimoto, M. Notch signaling in the mammalian respiratory system, specifically the trachea and lungs, in development, homeostasis, regeneration, and disease. Dev. Growth Differ. 2020, 62, 67–79. [Google Scholar] [CrossRef]
  30. Vanderbeck, A.; Maillard, I. Notch signaling at the crossroads of innate and adaptive immunity. J. Leukoc. Biol. 2021, 109, 535–548. [Google Scholar] [CrossRef]
  31. Jin, Y.; Wang, M.; Hu, H.; Huang, Q.; Chen, Y.; Wang, G. Overcoming stemness and chemoresistance in colorectal cancer through miR-195-5p-modulated inhibition of notch signaling. Int. J. Biol. Macromol. 2018, 117, 445–453. [Google Scholar] [CrossRef]
  32. Kukcinaviciute, E.; Jonusiene, V.; Sasnauskiene, A.; Dabkeviciene, D.; Eidenaite, E.; Laurinavicius, A. Significance of Notch and Wnt signaling for chemoresistance of colorectal cancer cells HCT116. J. Cell. Biochem. 2018, 119, 5913–5920. [Google Scholar] [CrossRef] [PubMed]
  33. Janghorban, M.; Xin, L.; Rosen, J.M.; Zhang, X.H.F. Notch signaling as a regulator of the tumor immune response: To target or not to target? Front. Immunol. 2018, 9, 1649. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, F.; Huang, C.; Long, J.; Zhao, Z.B.; Ma, H.Q.; Yao, X.Q.; Li, L.; Lian, Z.X. Notch signaling mutations increase intra-tumor chemokine expression and predict response to immunotherapy in colorectal cancer. BMC Cancer 2022, 22, 933. [Google Scholar] [CrossRef]
  35. Nandi, A.; Debnath, R.; Nayak, A.; To, T.K.J.; Thacker, G.; Reilly, M.; Gumber, S.; Karagounis, I.; Li, N.; Lengner, C.J.; et al. Dll1-Mediated Notch Signaling Drives Tumor Cell Cross-talk with Cancer-Associated Fibroblasts to Promote Radioresistance in Breast Cancer. Cancer Res. 2022, 82, 3718–3733. [Google Scholar] [CrossRef]
  36. Xie, Q.; Guo, H.; He, P.; Deng, H.; Gao, Y.; Dong, N.; Niu, W.; Liu, T.; Li, M.; Wang, S.; et al. Tspan5 promotes epithelial–mesenchymal transition and tumour metastasis of hepatocellular carcinoma by activating Notch signalling. Mol. Oncol. 2021, 15, 3184. [Google Scholar] [CrossRef] [PubMed]
  37. Choi, S.; Yu, J.; Park, A.; Dubon, M.J.; Do, J.; Kim, Y.; Nam, D.; Noh, J.; Park, K.S. BMP-4 enhances epithelial mesenchymal transition and cancer stem cell properties of breast cancer cells via Notch signaling. Sci. Rep. 2019, 9, 11724. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, L.; Sha, J.; Yang, G.; Huang, X.; Bo, J.; Huang, Y. Activation of Notch pathway is linked with epithelial-mesenchymal transition in prostate cancer cells. Cell Cycle 2017, 16, 999–1007. [Google Scholar] [CrossRef]
  39. Xie, Q.; Cheng, Z.; Chen, X.; Lobe, C.G.; Liu, J. The role of Notch signalling in ovarian angiogenesis. J. Ovarian Res. 2017, 10, 13. [Google Scholar] [CrossRef]
  40. Xiao, W.; Gao, Z.; Duan, Y.; Yuan, W.; Ke, Y. Notch signaling plays a crucial role in cancer stem-like cells maintaining stemness and mediating chemotaxis in renal cell carcinoma. J. Exp. Clin. Cancer Res. 2017, 36, 41. [Google Scholar] [CrossRef]
  41. Forghanifard, M.M.; Kasebi, P.; Abbaszadegan, M.R. SOX2/SALL4 stemness axis modulates Notch signaling genes to maintain self-renewal capacity of esophageal squamous cell carcinoma. Mol. Cell. Biochem. 2021, 476, 921–929. [Google Scholar] [CrossRef] [PubMed]
  42. Xiu, M.; Wang, Y.; Li, B.; Wang, X.; Xiao, F.; Chen, S.; Zhang, L.; Zhou, B.; Hua, F. The Role of Notch3 Signaling in Cancer Stemness and Chemoresistance: Molecular Mechanisms and Targeting Strategies. Front. Mol. Biosci. 2021, 8, 694141. [Google Scholar] [CrossRef] [PubMed]
  43. Kimura, S.; Seki, M.; Yoshida, K.; Shiraishi, Y.; Akiyama, M.; Koh, K.; Imamura, T.; Manabe, A.; Hayashi, Y.; Kobayashi, M.; et al. NOTCH1 pathway activating mutations and clonal evolution in pediatric T-cell acute lymphoblastic leukemia. Cancer Sci. 2019, 110, 784–794. [Google Scholar] [CrossRef]
  44. Takam Kamga, P.; Collo, G.D.; Resci, F.; Bazzoni, R.; Mercuri, A.; Quaglia, F.M.; Tanasi, I.; Delfino, P.; Visco, C.; Bonifacio, M.; et al. Notch Signaling Molecules as Prognostic Biomarkers for Acute Myeloid Leukemia. Cancers 2019, 11, 1958. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, C.; Berndt-Paetz, M.; Neuhaus, J. A comprehensive bioinformatics analysis of notch pathways in bladder cancer. Cancers 2021, 13, 3089. [Google Scholar] [CrossRef] [PubMed]
  46. Lin, Q.-Q.; Sun, J.-L.; Wang, F.; Zhang, H.-Z.; Zhou, G.; Xi, Q. Current understanding of adenoid cystic carcinoma in the gene expression and targeted therapy. Holist. Integr. Oncol. 2023, 2, 7. [Google Scholar] [CrossRef]
  47. Sorrentino, C.; Cuneo, A.; Roti, G. Therapeutic targeting of notch signaling pathway in hematological malignancies. Mediterr. J. Hematol. Infect. Dis. 2019, 11, e2019037. [Google Scholar]
  48. 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]
  49. Toribio, M.L.; González-García, S. Notch Partners in the Long Journey of T-ALL Pathogenesis. Int. J. Mol. Sci. 2023, 24, 1383. [Google Scholar] [CrossRef]
  50. Chimento, A.; D’amico, M.; Pezzi, V.; De Amicis, F. Notch Signaling in Breast Tumor Microenvironment as Mediator of Drug Resistance. Int. J. Mol. Sci. 2022, 23, 6296. [Google Scholar] [CrossRef]
  51. Nandi, A.; Chakrabarti, R. The many facets of Notch signaling in breast cancer: Toward overcoming therapeutic resistance. Genes Dev. 2020, 34, 1422–1438. [Google Scholar] [CrossRef] [PubMed]
  52. Edwards, A.; Brennan, K. Notch Signalling in Breast Development and Cancer. Front. Cell Dev. Biol. 2021, 9, 692173. [Google Scholar] [CrossRef] [PubMed]
  53. Yousefi, H.; Bahramy, A.; Zafari, N.; Delavar, M.R.; Nguyen, K.; Haghi, A.; Kandelouei, T.; Vittori, C.; Jazireian, P.; Maleki, S.; et al. Notch signaling pathway: A comprehensive prognostic and gene expression profile analysis in breast cancer. BMC Cancer 2022, 22, 1282. [Google Scholar] [CrossRef] [PubMed]
  54. Ding, L.; Bailey, M.H.; Porta-Pardo, E.; Thorsson, V.; Colaprico, A.; Bertrand, D.; Gibbs, D.L.; Weerasinghe, A.; Huang, K.L.; Tokheim, C.; et al. Perspective on Oncogenic Processes at the End of the Beginning of Cancer Genomics. Cell 2018, 173, 305–320. [Google Scholar] [CrossRef]
  55. Gao, J.; Aksoy, B.A.; Dogrusoz, U.; Dresdner, G.; Gross, B.; Sumer, S.O.; Sun, Y.; Jacobsen, A.; Sinha, R.; Larsson, E.; et al. Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Sci. Signal. 2013, 6, pl1. [Google Scholar] [CrossRef]
  56. Cerami, E.; Gao, J.; Dogrusoz, U.; Gross, B.E.; Sumer, S.O.; Aksoy, B.A.; Jacobsen, A.; Byrne, C.J.; Heuer, M.L.; Larsson, E.; et al. The cBio Cancer Genomics Portal: An open platform for exploring multidimensional cancer genomics data. Cancer Discov. 2012, 2, 401–404. [Google Scholar] [CrossRef]
  57. Zhang, M.; Singh, R.; Peng, S.; Mazumdar, T.; Sambandam, V.; Shen, L.; Tong, P.; Li, L.; Kalu, N.N.; Pickering, C.R.; et al. Mutations of the LIM protein AJUBA mediate sensitivity of head and neck squamous cell carcinoma to treatment with cell-cycle inhibitors. Cancer Lett. 2017, 392, 71–82. [Google Scholar] [CrossRef]
  58. Yokobori, T.; Mimori, K.; Iwatsuki, M.; Ishii, H.; Tanaka, F.; Sato, T.; Toh, H.; Sudo, T.; Iwaya, T.; Tanaka, Y.; et al. Copy number loss of FBXW7 is related to gene expression and poor prognosis in esophageal squamous cell carcinoma. Int. J. Oncol. 2012, 41, 253–259. [Google Scholar]
  59. Schleicher, K.; Schramek, D. AJUBA: A regulator of epidermal homeostasis and cancer. Exp. Dermatol. 2021, 30, 546–559. [Google Scholar] [CrossRef]
  60. Luo, Z.; Mu, L.; Zheng, Y.; Shen, W.; Li, J.; Xu, L.; Zhong, B.; Liu, Y.; Zhou, Y. NUMB enhances Notch signaling by repressing ubiquitination of NOTCH1 intracellular domain. J. Mol. Cell Biol. 2020, 12, 345–358. [Google Scholar] [CrossRef]
  61. Kar, R.; Jha, S.K.; Ojha, S.; Sharma, A.; Dholpuria, S.; Raju, V.S.R.; Prasher, P.; Chellappan, D.K.; Gupta, G.; Kumar Singh, S.; et al. The FBXW7-NOTCH interactome: A ubiquitin proteasomal system-induced crosstalk modulating oncogenic transformation in human tissues. Cancer Rep. 2021, 4, e1369. [Google Scholar] [CrossRef] [PubMed]
  62. Grilli, G.; Hermida-Prado, F.; Álvarez-Fernández, M.; Allonca, E.; Álvarez-González, M.; Astudillo, A.; Moreno-Bueno, G.; Cano, A.; García-Pedrero, J.M.; Rodrigo, J.P. Impact of notch signaling on the prognosis of patients with head and neck squamous cell carcinoma. Oral Oncol. 2020, 110, 105003. [Google Scholar] [CrossRef] [PubMed]
  63. Di Nardo, L.; Pellegrini, C.; Di Stefani, A.; Del Regno, L.; Sollena, P.; Piccerillo, A.; Longo, C.; Garbe, C.; Fargnoli, M.C.; Peris, K. Molecular genetics of cutaneous squamous cell carcinoma: Perspective for treatment strategies. J. Eur. Acad. Dermatol. Venereol. 2020, 34, 932–941. [Google Scholar] [CrossRef]
  64. 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]
  65. Schwaederle, M.; Elkin, S.K.; Tomson, B.N.; Carter, J.L.; Kurzrock, R. Squamousness: Next-generation sequencing reveals shared molecular features across squamous tumor types. Cell Cycle 2015, 14, 2355–2361. [Google Scholar] [CrossRef] [PubMed]
  66. Nyman, P.E.; Buehler, D.; Lambert, P.F. Loss of function of canonical Notch signaling drives head and neck carcinogenesis. Clin. Cancer Res. 2018, 24, 6308–6318. [Google Scholar] [CrossRef]
  67. Kałafut, J.; Czerwonka, A.; Anameriç, A.; Przybyszewska-Podstawka, A.; Misiorek, J.O.; Rivero-Müller, A.; Nees, M. Shooting at Moving and Hidden Targets-Tumour Cell Plasticity and the Notch Signalling Pathway in Head and Neck Squamous Cell Carcinomas. Cancers 2021, 13, 6219. [Google Scholar] [CrossRef]
  68. Shah, P.A.; Huang, C.; Li, Q.; Kazi, S.A.; Byers, L.A.; Wang, J.; Johnson, F.M.; Frederick, M.J. NOTCH1 Signaling in Head and Neck Squamous Cell Carcinoma. Cells 2020, 9, 2677. [Google Scholar] [CrossRef]
  69. Nowell, C.S.; Radtke, F. Notch as a tumour suppressor. Nat. Rev. Cancer 2017, 17, 145–159. [Google Scholar] [CrossRef]
  70. Zhang, M.; Biswas, S.; Qin, X.; Gong, W.; Deng, W.; Yu, H. Does Notch play a tumor suppressor role across diverse squamous cell carcinomas? Cancer Med. 2016, 5, 2048–2060. [Google Scholar] [CrossRef]
  71. 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. Engl. J. Med. 2013, 369, 341–350. [Google Scholar] [CrossRef] [PubMed]
  72. Huang, C.; Chen, L.; Savage, S.R.; Eguez, R.V.; Dou, Y.; Li, Y.; da Veiga Leprevost, F.; Jaehnig, E.J.; Lei, J.T.; Wen, B.; et al. Proteogenomic insights into the biology and treatment of HPV-negative head and neck squamous cell carcinoma. Cancer Cell 2021, 39, 361–379. [Google Scholar] [CrossRef] [PubMed]
  73. Ali, J.; Sabiha, B.; Jan, H.U.; Haider, S.A.; Khan, A.A.; Ali, S.S. Genetic etiology of oral cancer. Oral Oncol. 2017, 70, 23–28. [Google Scholar] [CrossRef] [PubMed]
  74. Pickering, C.R.; Zhang, J.; Yoo, S.Y.; Bengtsson, L.; Moorthy, S.; Neskey, D.M.; Zhao, M.; Ortega Alves, M.V.; Chang, K.; Drummond, J.; et al. Integrative genomic characterization of oral squamous cell carcinoma identifies frequent somatic drivers. Cancer Discov. 2013, 3, 770–781. [Google Scholar] [CrossRef]
  75. Hoadley, K.A.; Yau, C.; Hinoue, T.; Wolf, D.M.; Lazar, A.J.; Drill, E.; Shen, R.; Taylor, A.M.; Cherniack, A.D.; Thorsson, V.; et al. Cell-of-Origin Patterns Dominate the Molecular Classification of 10,000 Tumors from 33 Types of Cancer. Cell 2018, 173, 291–304. [Google Scholar] [CrossRef]
  76. Sanchez-Vega, F.; Mina, M.; Armenia, J.; Chatila, W.K.; Luna, A.; La, K.C.; Dimitriadoy, S.; Liu, D.L.; Kantheti, H.S.; Saghafinia, S.; et al. Oncogenic Signaling Pathways in The Cancer Genome Atlas. Cell 2018, 173, 321–337. [Google Scholar] [CrossRef]
  77. 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]
  78. Schmidl, B.; Siegl, M.; Boxberg, M.; Stögbauer, F.; Jira, D.; Winter, C.; Stark, L.; Pickhard, A.; Wollenberg, B.; Wirth, M. NOTCH1 Intracellular Domain and the Tumor Microenvironment as Prognostic Markers in HNSCC. Cancers 2022, 14, 1080. [Google Scholar] [CrossRef]
  79. Sun, W.; Gaykalova, D.A.; Ochs, M.F.; Mambo, E.; Arnaoutakis, D.; Liu, Y.; Loyo, M.; Agrawal, N.; Howard, J.; Li, R.; et al. Activation of the NOTCH pathway in head and neck cancer. Cancer Res. 2014, 74, 1091–1104. [Google Scholar] [CrossRef]
  80. Porcheri, C.; Mitsiadis, T.A. Notch in Head and Neck Cancer. Adv. Exp. Med. Biol. 2021, 1287, 81–103. [Google Scholar]
  81. Porcheri, C.; Meisel, C.T.; Mitsiadis, T. Multifactorial Contribution of Notch Signaling in Head and Neck Squamous Cell Carcinoma. Int. J. Mol. Sci. 2019, 20, 1520. [Google Scholar] [CrossRef] [PubMed]
  82. Weaver, A.N.; Benjamin Burch, M.; Cooper, T.S.; Manna, D.L.D.; Wei, S.; Ojesina, A.I.; Rosenthal, E.L.; Yang, E.S. Notch signaling activation is associated with patient mortality and increased FGF1-mediated invasion in squamous cell carcinoma of the oral cavity. Mol. Cancer Res. 2016, 14, 883–891. [Google Scholar] [CrossRef] [PubMed]
  83. Loganathan, S.K.; Schleicher, K.; Malik, A.; Quevedo, R.; Langille, E.; Teng, K.; Oh, R.H.; Rathod, B.; Tsai, R.; Samavarchi-Tehrani, P.; et al. Rare driver mutations in head and neck squamous cell carcinomas converge on NOTCH signaling. Science 2020, 367, 1264–1269. [Google Scholar] [CrossRef] [PubMed]
  84. Kitamoto, T.; Takahashi, K.; Takimoto, H.; Tomizuka, K.; Hayasaka, M.; Tabira, T.; Hanaoka, K. Functional redundancy of the Notch gene family during mouse embryogenesis: Analysis of Notch gene expression in Notch3-deficient mice. Biochem. Biophys. Res. Commun. 2005, 331, 1154–1162. [Google Scholar] [CrossRef] [PubMed]
  85. 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]
  86. Riccio, O.; van Gijn, M.E.; Bezdek, A.C.; Pellegrinet, L.; van Es, J.H.; Zimber-Strobl, U.; Strobl, L.J.; Honjo, T.; Clevers, H.; Radtke, F. Loss of intestinal crypt progenitor cells owing to inactivation of both Notch1 and Notch2 is accompanied by derepression of CDK inhibitors p27Kip1 and p57Kip2. EMBO Rep. 2008, 9, 377–383. [Google Scholar] [CrossRef]
  87. Hatano, K.; Saigo, C.; Kito, Y.; Shibata, T.; Takeuchi, T. Overexpression of JAG2 is related to poor outcomes in oral squamous cell carcinoma. Clin. Exp. Dent. Res. 2020, 6, 174–180. [Google Scholar] [CrossRef]
  88. Lin, J.T.; Chen, M.K.; Yeh, K.T.; Chang, C.S.; Chang, T.H.; Lin, C.Y.; Wu, Y.C.; Su, B.W.; Lee, K.D.; Chang, P.J. Association of high levels of Jagged-1 and Notch-1 expression with poor prognosis in head and neck cancer. Ann. Surg. Oncol. 2010, 17, 2976–2983. [Google Scholar] [CrossRef]
  89. Zhang, T.; Liang, L.; Liu, X.; Wu, J.N.; Chen, J.; Su, K.; Zheng, Q.; Huang, H.; Liao, G. qing TGFβ1-Smad3-Jagged1-Notch1-Slug signaling pathway takes part in tumorigenesis and progress of tongue squamous cell carcinoma. J. Oral Pathol. Med. 2016, 45, 486–493. [Google Scholar] [CrossRef]
  90. Parmigiani, E.; Taylor, V.; Giachino, C. Oncogenic and Tumor-Suppressive Functions of NOTCH Signaling in Glioma. Cells 2020, 9, 2304. [Google Scholar] [CrossRef]
  91. Marignol, L. Notch signalling: The true driver of small cell lung cancer? Transl. Cancer Res. 2017, 6, 1191–1196. [Google Scholar] [CrossRef]
  92. Leonetti, A.; Facchinetti, F.; Minari, R.; Cortellini, A.; Rolfo, C.D.; Giovannetti, E.; Tiseo, M. Notch pathway in small-cell lung cancer: From preclinical evidence to therapeutic challenges. Cell. Oncol. 2019, 42, 261–273. [Google Scholar] [CrossRef] [PubMed]
  93. Shah, P.A.; Sambandam, V.; Fernandez, A.M.; Zhao, H.; Mazumdar, T.; Shen, L.; Wang, Q.; Ahmed, K.M.; Ghosh, S.; Frederick, M.J.; et al. Sustained Aurora Kinase B Expression Confers Resistance to PI3K Inhibition in Head and Neck Squamous Cell Carcinoma. Cancer Res. 2022, 82, 4444–4456. [Google Scholar] [CrossRef] [PubMed]
  94. Nolin, E.; Gans, S.; Llamas, L.; Bandyopadhyay, S.; Brittain, S.M.; Bernasconi-Elias, P.; Carter, K.P.; Loureiro, J.J.; Thomas, J.R.; Schirle, M.; et al. Discovery of a ZIP7 inhibitor from a Notch pathway screen. Nat. Chem. Biol. 2019, 15, 179–188. [Google Scholar] [CrossRef] [PubMed]
  95. Roti, G.; Carlton, A.; Ross, K.N.; Markstein, M.; Pajcini, K.; Su, A.H.; Perrimon, N.; Pear, W.S.; Kung, A.L.; Blacklow, S.C.; et al. Complementary genomic screens identify SERCA as a therapeutic target in NOTCH1 mutated cancer. Cancer Cell 2013, 23, 390–405. [Google Scholar] [CrossRef] [PubMed]
  96. Isaacs, J.T.; Brennen, W.N.; Christensen, S.B.; Denmeade, S.R. Mipsagargin: The beginning—Not the end—Of thapsigargin prodrug-based cancer therapeutics. Molecules 2021, 26, 7469. [Google Scholar] [CrossRef]
  97. De Ford, C.; Heidersdorf, B.; Haun, F.; Murillo, R.; Friedrich, T.; Borner, C.; Merfort, I. The clerodane diterpene casearin J induces apoptosis of T-all cells through SERCA inhibition, oxidative stress, and interference with notch1 signaling. Cell Death Dis. 2016, 7, e2070. [Google Scholar] [CrossRef]
  98. Marchesini, M.; Gherli, A.; Montanaro, A.; Patrizi, L.; Sorrentino, C.; Pagliaro, L.; Rompietti, C.; Kitara, S.; Heit, S.; Olesen, C.E.; et al. Blockade of Oncogenic NOTCH1 with the SERCA Inhibitor CAD204520 in T Cell Acute Lymphoblastic Leukemia. Cell Chem. Biol. 2020, 27, 678–697. [Google Scholar] [CrossRef]
  99. Pagliaro, L.; Marchesini, M.; Roti, G. Targeting oncogenic Notch signaling with SERCA inhibitors. J. Hematol. Oncol. 2021, 14, 8. [Google Scholar] [CrossRef]
  100. Pagliaro, L.; Sorrentino, C.; Roti, G. Targeting Notch Trafficking and Processing in Cancers. Cells 2020, 9, 2212. [Google Scholar] [CrossRef]
  101. Baldoni, S.; Del Papa, B.; Dorillo, E.; Aureli, P.; De Falco, F.; Rompietti, C.; Sorcini, D.; Varasano, E.; Cecchini, D.; Zei, T.; et al. Bepridil exhibits anti-leukemic activity associated with NOTCH1 pathway inhibition in chronic lymphocytic leukemia. Int. J. Cancer 2018, 143, 958–970. [Google Scholar] [CrossRef]
  102. Krämer, A.; Mentrup, T.; Kleizen, B.; Rivera-Milla, E.; Reichenbach, D.; Enzensperger, C.; Nohl, R.; Täuscher, E.; Görls, H.; Ploubidou, A.; et al. Small molecules intercept Notch signaling and the early secretory pathway. Nat. Chem. Biol. 2013, 9, 731–738. [Google Scholar] [CrossRef]
  103. Zhang, M.; Han, Y.; Zheng, Y.; Zhang, Y.; Zhao, X.; Gao, Z.; Liu, X. ZEB1-activated LINC01123 accelerates the malignancy in lung adenocarcinoma through NOTCH signaling pathway. Cell Death Dis. 2020, 11, 981. [Google Scholar] [CrossRef]
  104. Lu, Z.; Ren, Y.; Zhang, M.; Fan, T.; Wang, Y.; Zhao, Q.; Liu, H.M.; Zhao, W.; Hou, G. FLI-06 suppresses proliferation, induces apoptosis and cell cycle arrest by targeting LSD1 and Notch pathway in esophageal squamous cell carcinoma cells. Biomed. Pharmacother. 2018, 107, 1370–1376. [Google Scholar] [CrossRef]
  105. Gan, R.H.; Lin, L.S.; Xie, J.; Huang, L.; Ding, L.C.; Su, B.H.; Peng, X.E.; Zheng, D.L.; Lu, Y.G. FLI-06 Intercepts Notch Signaling and Suppresses the Proliferation and Self-renewal of Tongue Cancer Cells. OncoTargets Ther. 2019, 12, 7663–7674. [Google Scholar] [CrossRef] [PubMed]
  106. Hounjet, J.; Habets, R.; Schaaf, M.B.; Hendrickx, T.C.; Barbeau, L.M.O.; Yahyanejad, S.; Rouschop, K.M.; Groot, A.J.; Vooijs, M. The anti-malarial drug chloroquine sensitizes oncogenic NOTCH1 driven human T-ALL to γ-secretase inhibition. Oncogene 2019, 38, 5457–5468. [Google Scholar] [CrossRef] [PubMed]
  107. Kobia, F.; Duchi, S.; Deflorian, G.; Vaccari, T. Pharmacologic inhibition of vacuolar H+ ATPase reduces physiologic and oncogenic Notch signaling. Mol. Oncol. 2014, 8, 207–220. [Google Scholar] [CrossRef] [PubMed]
  108. Schneider, M.; Kumar, V.; Nordstrøm, L.U.; Feng, L.; Takeuchi, H.; Hao, H.; Luca, V.C.; Garcia, K.C.; Stanley, P.; Wu, P.; et al. Inhibition of Delta-induced Notch signaling using fucose analogs. Nat. Chem. Biol. 2018, 14, 65–71. [Google Scholar] [CrossRef] [PubMed]
  109. Wolfe, M.S. Structure and Function of the γ-Secretase Complex. Biochemistry 2019, 58, 2953–2966. [Google Scholar] [CrossRef]
  110. Bai, X.C.; Rajendra, E.; Yang, G.; Shi, Y.; Scheres, S.H.W. Sampling the conformational space of the catalytic subunit of human g-secretase. Elife 2015, 4, e11182. [Google Scholar] [CrossRef]
  111. Hur, J.Y. γ-Secretase in Alzheimer’s disease. Exp. Mol. Med. 2022, 54, 433–446. [Google Scholar] [CrossRef] [PubMed]
  112. Luo, J.E.; Li, Y.M. Turning the tide on Alzheimer’s disease: Modulation of γ-secretase. Cell Biosci. 2022, 12, 2. [Google Scholar] [CrossRef] [PubMed]
  113. Deangelo, D.J.; Stone, R.M.; Silverman, L.B.; Stock, W.; Attar, E.C.; Fearen, I.; Dallob, A.; Matthews, C.; Stone, J.; Freedman, S.J.; et al. A phase I clinical trial of the notch inhibitor MK-0752 in patients with T-cell acute lymphoblastic leukemia/lymphoma (T-ALL) and other leukemias. J. Clin. Oncol. 2006, 24, 6585. [Google Scholar] [CrossRef]
  114. López-Nieva, P.; González-Sánchez, L.; Cobos-Fernández, M.Á.; Córdoba, R.; Santos, J.; Fernández-Piqueras, J. More Insights on the Use of γ-Secretase Inhibitors in Cancer Treatment. Oncologist 2021, 26, e298–e305. [Google Scholar] [CrossRef]
  115. Pine, S.R. Rethinking Gamma-secretase Inhibitors for Treatment of Non–small-Cell Lung Cancer: Is notch the target? Clin. Cancer Res. 2018, 24, 6136–6141. [Google Scholar] [CrossRef] [PubMed]
  116. Ran, Y.; Hossain, F.; Pannuti, A.; Lessard, C.B.; Ladd, G.Z.; Jung, J.I.; Minter, L.M.; Osborne, B.A.; Miele, L.; Golde, T.E. γ-Secretase inhibitors in cancer clinical trials are pharmacologically and functionally distinct. EMBO Mol. Med. 2017, 9, 950–966. [Google Scholar] [CrossRef]
  117. 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 γ-secretase inhibitor PF-03084014 in combination with glucocorticoids in T-cell acute lymphoblastic leukemia. Mol. Cancer Ther. 2012, 11, 1565–1575. [Google Scholar] [CrossRef]
  118. Collins, M.; Michot, J.M.; Bellanger, C.; Mussini, C.; Benhadji, K.; Massard, C.; Carbonnel, F. Notch inhibitors induce diarrhea, hypercrinia and secretory cell Metaplasia in the human colon. EXCLI J. 2021, 20, 819–827. [Google Scholar]
  119. Sancho, R.; Cremona, C.A.; Behrens, A. Stem cell and progenitor fate in the mammalian intestine: Notch and lateral inhibition in homeostasis and disease. EMBO Rep. 2015, 16, 571–581. [Google Scholar] [CrossRef]
  120. Chan, D.; Kaplan, J.; Gordon, G.; Desai, J. Activity of the gamma secretase inhibitor al101 in desmoid tumors: A case report of 2 adult cases. Curr. Oncol. 2021, 28, 3659–3667. [Google Scholar] [CrossRef]
  121. Federman, N. Molecular pathogenesis of desmoid tumor and the role of γ-secretase inhibition. NPJ Precis. Oncol. 2022, 6, 62. [Google Scholar] [CrossRef] [PubMed]
  122. Gounder, M.; Ratan, R.; Alcindor, T.; Schöffski, P.; van der Graaf, W.T.; Wilky, B.A.; Riedel, R.F.; Lim, A.; Smith, L.M.; Moody, S.; et al. Nirogacestat, a γ-Secretase Inhibitor for Desmoid Tumors. N. Engl. J. Med. 2023, 388, 898–912. [Google Scholar] [CrossRef] [PubMed]
  123. Even, C.; Lassen, U.; Merchan, J.; Le Tourneau, C.; Soria, J.C.; Ferte, C.; Ricci, F.; Diener, J.T.; Yuen, E.; Smith, C.; et al. Safety and clinical activity of the Notch inhibitor, crenigacestat (LY3039478), in an open-label phase I trial expansion cohort of advanced or metastatic adenoid cystic carcinoma. Investig. New Drugs 2020, 38, 402–409. [Google Scholar] [CrossRef] [PubMed]
  124. Massard, C.; Cassier, P.A.; Azaro, A.; Anderson, B.; Yuen, E.; Yu, D.; Oakley, G.; Benhadji, K.A.; Pant, S. A phase 1b study of crenigacestat (LY3039478) in combination with gemcitabine and cisplatin or gemcitabine and carboplatin in patients with advanced or metastatic solid tumors. Cancer Chemother. Pharmacol. 2022, 90, 335–344. [Google Scholar] [CrossRef]
  125. Doi, T.; Tajimi, M.; Mori, J.; Asou, H.; Inoue, K.; Benhadji, K.A.; Naito, Y. A phase 1 study of crenigacestat (LY3039478), the Notch inhibitor, in Japanese patients with advanced solid tumors. Investig. New Drugs 2021, 39, 469–476. [Google Scholar] [CrossRef] [PubMed]
  126. Borthakur, G.; Martinelli, G.; Raffoux, E.; Chevallier, P.; Chromik, J.; Lithio, A.; Smith, C.L.; Yuen, E.; Oakley, G.J.; Benhadji, K.A.; et al. Phase 1 study to evaluate Crenigacestat (LY3039478) in combination with dexamethasone in patients with T-cell acute lymphoblastic leukemia and lymphoma. Cancer 2021, 127, 372–380. [Google Scholar] [CrossRef]
  127. McCaw, T.R.; Inga, E.; Chen, H.; Jaskula-Sztul, R.; Dudeja, V.; Bibb, J.A.; Ren, B.; Rose, J.B. Gamma Secretase Inhibitors in Cancer: A Current Perspective on Clinical Performance. Oncologist 2021, 26, e608–e621. [Google Scholar] [CrossRef]
  128. Tagami, S.; Yanagida, K.; Kodama, T.S.; Takami, M.; Mizuta, N.; Oyama, H.; Nishitomi, K.; Chiu, Y.W.; Okamoto, T.; Ikeuchi, T.; et al. Semagacestat Is a Pseudo-Inhibitor of γ-Secretase. Cell Rep. 2017, 21, 259–273. [Google Scholar] [CrossRef]
  129. Gu, K.; Li, Q.; Lin, H.; Zhu, J.; Mo, J.; He, S.; Lu, X.; Jiang, X.; Sun, H. Gamma secretase inhibitors: A patent review (2013–2015). Expert Opin. Ther. Pat. 2017, 27, 851–866. [Google Scholar] [CrossRef]
  130. Zhu, R.R.; Chen, Q.; Liu, Z.B.; Ruan, H.G.; Wu, Q.C.; Zhou, X. liang Inhibition of the Notch1 pathway induces peripartum cardiomyopathy. J. Cell. Mol. Med. 2020, 24, 7907–7914. [Google Scholar] [CrossRef]
  131. Sajadimajd, S.; Bahrami, G.; Mohammadi, B.; Madani, S.H. Notch signaling-induced cyclin d1 in diabetes ameliorating effects of the isolated polysaccharide from Rosa canina: In vitro and in vivo studies. Cell Biochem. Funct. 2022, 40, 935–945. [Google Scholar] [CrossRef]
  132. Kim, H.K.; Lee, D.W.; Kim, E.; Jeong, I.; Kim, S.; Kim, B.J.; Park, H.C. Notch signaling controls oligodendrocyte regeneration in the injured telencephalon of adult zebrafish. Exp. Neurobiol. 2020, 29, 417–424. [Google Scholar] [CrossRef]
  133. Ohuchi, K.; Funato, M.; Yoshino, Y.; Ando, S.; Inagaki, S.; Sato, A.; Kawase, C.; Seki, J.; Saito, T.; Nishio, H.; et al. Notch Signaling Mediates Astrocyte Abnormality in Spinal Muscular Atrophy Model Systems. Sci. Rep. 2019, 9, 3701. [Google Scholar] [CrossRef] [PubMed]
  134. Aguayo-Ortiz, R.; Guzmán-Ocampo, D.C.; Dominguez, L. Toward the Characterization of DAPT Interactions with γ-Secretase. ChemMedChem 2019, 14, 1005–1010. [Google Scholar] [CrossRef]
  135. Dai, G.; Deng, S.; Guo, W.; Yu, L.; Yang, J.; Zhou, S.; Gao, T. Notch pathway inhibition using DAPT, a γ-secretase inhibitor (GSI), enhances the antitumor effect of cisplatin in resistant osteosarcoma. Mol. Carcinog. 2019, 58, 3–18. [Google Scholar] [CrossRef] [PubMed]
  136. Pindiprolu, S.K.S.S.; Krishnamurthy, P.T.; Dev, C.; Chintamaneni, P.K. DR5 antibody conjugated lipid-based nanocarriers of gamma-secretase inhibitor for the treatment of triple negative breast cancer. Chem. Phys. Lipids 2021, 235, 105033. [Google Scholar] [CrossRef] [PubMed]
  137. Ghanbari-Movahed, M.; Shiri Varnamkhasti, B.; Shourian, M. Inhibiting Notch activity in breast cancer stem cells by functionalized gold nanoparticles with gamma-secretase inhibitor DAPT and vitamin C. Chem. Pap. 2022, 76, 1157–1170. [Google Scholar] [CrossRef]
  138. Wan, X.; Liu, C.; Lin, Y.; Fu, J.; Lu, G.; Lu, Z. pH sensitive peptide functionalized nanoparticles for co-delivery of erlotinib and DAPT to restrict the progress of triple negative breast cancer. Drug Deliv. 2019, 26, 470–480. [Google Scholar] [CrossRef]
  139. Lee, C.; Kim, M.; Park, C.; Jo, W.; Seo, J.K.; Kim, S.; Oh, J.; Kim, C.S.; Ryu, H.S.; Lee, K.H.; et al. Epigenetic regulation of Neuregulin 1 promotes breast cancer progression associated to hyperglycemia. Nat. Commun. 2023, 14, 439. [Google Scholar] [CrossRef]
  140. Bocchicchio, S.; Tesone, M.; Irusta, G. Convergence of Wnt and Notch signaling controls ovarian cancer cell survival. J. Cell. Physiol. 2019, 234, 22130–22143. [Google Scholar] [CrossRef]
  141. He, G.; Mu, T.; Yuan, Y.; Yang, W.; Zhang, Y.; Chen, Q.; Bian, M.; Pan, Y.; Xiang, Q.; Chen, Z.; et al. Effects of notch signaling pathway in cervical cancer by curcumin mediated photodynamic therapy and its possible mechanisms in vitro and in vivo. J. Cancer 2019, 10, 4114–4122. [Google Scholar] [CrossRef] [PubMed]
  142. Sun, T.; Zhang, D.; Wang, Z.; Zhao, B.; Li, Y.; Sun, X.; Liu, J.; Wang, X.; Sheng, J. Inhibition of the notch signaling pathway overcomes resistance of cervical cancer cells to paclitaxel through retardation of the epithelial–mesenchymal transition process. Environ. Toxicol. 2021, 36, 1758–1764. [Google Scholar] [CrossRef]
  143. Rice, M.A.; Hsu, E.C.; Aslan, M.; Ghoochani, A.; Su, A.; Stoyanova, T. Loss of Notch1 activity inhibits prostate cancer growth and metastasis and sensitizes prostate cancer cells to antiandrogen therapies. Mol. Cancer Ther. 2019, 18, 1230–1242. [Google Scholar] [CrossRef] [PubMed]
  144. Peng, X.; Zhou, J.; Li, B.; Zhang, T.; Zuo, Y.; Gu, X. Notch1 and PI3K/Akt signaling blockers DAPT and LY294002 coordinately inhibit metastasis of gastric cancer through mutual enhancement. Cancer Chemother. Pharmacol. 2020, 85, 309–320. [Google Scholar] [CrossRef] [PubMed]
  145. He, Y.; He, P.; Lu, S.; Dong, W. KIFC3 Regulates the progression and metastasis of gastric cancer via Notch1 pathway. Dig. Liver Dis. 2023, 55, 1270–1279. [Google Scholar] [CrossRef] [PubMed]
  146. Clark, A.G.; Bertrand, F.E.; Sigounas, G. A potential requirement for Smad3 phosphorylation in Notch-mediated EMT in colon cancer. Adv. Biol. Regul. 2023, 88, 100957. [Google Scholar] [CrossRef]
  147. Liu, F.; Chu, H.X.; Han, J.S.; Sun, X.; Chen, J.; Qiu, X.L.; Zheng, X.H.; Jia, B.; Zhao, J.J. Inhibitory effect of the Notch pathway-inhibitor DAPT on invasion and metastasis of tongue cancer via lncRNA-KAT14 regulation. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 189–199. [Google Scholar]
  148. Liu, S.; Dou, L.; Miao, M.; Man, X.; Wei, B.; Jiang, Z.; Ouyang, Y.; Ozaki, T.; Yu, M.; Zhu, Y. HES1-mediated down-regulation of miR-138 sustains NOTCH1 activation and promotes proliferation and invasion in renal cell carcinoma. J. Exp. Clin. Cancer Res. 2023, 42, 72. [Google Scholar] [CrossRef]
  149. Lou, W.; Gao, K.; Xu, C.; Li, Q. Bromodomain-containing protein 9 is a prognostic biomarker associated with immune infiltrates and promotes tumor malignancy through activating notch signaling pathway in negative HIF-2α clear cell renal cell carcinoma. IUBMB Life 2021, 73, 1334–1347. [Google Scholar] [CrossRef]
  150. Wu, Q.; Guo, J.; Liu, Y.; Zheng, Q.; Li, X.; Wu, C.; Fang, D.; Chen, X.; Ma, L.; Xu, P.; et al. YAP drives fate conversion and chemoresistance of small cell lung cancer. Sci. Adv. 2021, 7, eabg1850. [Google Scholar] [CrossRef]
  151. Mancarella, S.; Serino, G.; Coletta, S.; Armentano, R.; Dituri, F.; Ardito, F.; Ruzzenente, A.; Fabregat, I.; Giannelli, G. The Tumor Microenvironment Drives Intrahepatic Cholangiocarcinoma Progression. Int. J. Mol. Sci. 2022, 23, 4187. [Google Scholar] [CrossRef] [PubMed]
  152. Mancarella, S.; Serino, G.; Dituri, F.; Cigliano, A.; Ribback, S.; Wang, J.; Chen, X.; Calvisi, D.F.; Giannelli, G. Crenigacestat, a selective NOTCH1 inhibitor, reduces intrahepatic cholangiocarcinoma progression by blocking VEGFA/DLL4/MMP13 axis. Cell Death Differ. 2020, 27, 2330–2343. [Google Scholar] [CrossRef] [PubMed]
  153. Danielpour, D.; Corum, S.; Leahy, P.; Bangalore, A. Jagged-1 is induced by mTOR inhibitors in renal cancer cells through an Akt/ALK5/Smad4-dependent mechanism. Curr. Res. Pharmacol. Drug Discov. 2022, 3, 100117. [Google Scholar] [CrossRef] [PubMed]
  154. Low, H.Y.; Lee, Y.C.; Lee, Y.J.; Wang, H.L.; Chen, Y.I.; Chien, P.J.; Li, S.T.; Chang, W.W. Reciprocal regulation between indoleamine 2,3-dioxigenase 1 and notch1 involved in radiation response of cervical cancer stem cells. Cancers 2020, 12, 1547. [Google Scholar] [CrossRef] [PubMed]
  155. Giordano, F.; Montalto, F.I.; Panno, M.L.; Andò, S.; De Amicis, F. A Notch inhibitor plus Resveratrol induced blockade of autophagy drives glioblastoma cell death by promoting a switch to apoptosis. Am. J. Cancer Res. 2021, 11, 5933–5950. [Google Scholar]
  156. Bill, M.; Pathmanathan, A.; Karunasiri, M.; Shen, C.; Burke, M.H.; Ranganathan, P.; Papaioannou, D.; Zitzer, N.C.; Snyder, K.; Larocco, A.; et al. EGFL7 antagonizes notch signaling and represents a novel therapeutic target in acute myeloid leukemia. Clin. Cancer Res. 2020, 26, 669–678. [Google Scholar] [CrossRef]
  157. Li, T.; Xu, X.H.; Guo, X.; Yuan, T.; Tang, Z.H.; Jiang, X.M.; Xu, Y.L.; Zhang, L.L.; Chen, X.; Zhu, H.; et al. Activation of notch 3/c-MYC/CHOP axis regulates apoptosis and promotes sensitivity of lung cancer cells to mTOR inhibitor everolimus. Biochem. Pharmacol. 2020, 175, 113921. [Google Scholar] [CrossRef]
  158. Ghanbari-Movahed, M.; Ghanbari-Movahed, Z.; Momtaz, S.; Kilpatrick, K.L.; Farzaei, M.H.; Bishayee, A. Unlocking the secrets of cancer stem cells with -Secretase inhibitors: A novel anticancer strategy. Molecules 2021, 26, 972. [Google Scholar] [CrossRef]
  159. Pathak, Y.; Camps, I.; Mishra, A.; Tripathi, V. Targeting notch signaling pathway in breast cancer stem cells through drug repurposing approach. Mol. Divers. 2022. [Google Scholar] [CrossRef]
  160. Singh, A.K.; Prajapati, K.S.; Kumar, S. Hesperidin potentially interacts with the catalytic site of gamma-secretase and modifies notch sensitive genes and cancer stemness marker expression in colon cancer cells and colonosphere. J. Biomol. Struct. Dyn. 2022, 41, 8432–8444. [Google Scholar] [CrossRef]
  161. Subramaniam, D.; Ponnurangam, S.; Ramamoorthy, P.; Standing, D.; Battafarano, R.J.; Anant, S.; Sharma, P. Curcumin induces cell death in esophageal cancer cells through modulating Notch signaling. PLoS ONE 2012, 7, e30590. [Google Scholar] [CrossRef]
  162. Yang, X.; Zhang, Y.; Huang, Y.; Wang, Y.; Qi, X.; Su, T.; Lu, L. Evodiamine suppresses Notch3 signaling in lung tumorigenesis via direct binding to γ-secretases. Phytomedicine 2020, 68, 153176. [Google Scholar] [CrossRef]
  163. Nie, P.; Kalidindi, T.; Nagle, V.L.; Wu, X.; Li, T.; Liao, G.P.; Frost, G.; Henry, K.E.; Punzalan, B.; Carter, L.M.; et al. Imaging of Cancer γ-Secretase activity using an Inhibitor-Based PET Probe. Clin. Cancer Res. 2021, 27, 6145–6155. [Google Scholar] [CrossRef]
  164. Yang, G.; Zhou, R.; Guo, X.; Yan, C.; Lei, J.; Shi, Y. Structural basis of γ-secretase inhibition and modulation by small molecule drugs. Cell 2021, 184, 521–533. [Google Scholar] [CrossRef] [PubMed]
  165. Chen, H.; Yu, T.; Lin, L.; Xing, L.; Cho, S.F.; Wen, K.; Aardalen, K.; Oka, A.; Lam, J.; Daley, M.; et al. γ-secretase inhibitors augment efficacy of BCMA-targeting bispecific antibodies against multiple myeloma cells without impairing T-cell activation and differentiation. Blood Cancer J. 2022, 12, 118. [Google Scholar] [CrossRef] [PubMed]
  166. Ambrogio, C.; Gómez-López, G.; Falcone, M.; Vidal, A.; Nadal, E.; Crosetto, N.; Blasco, R.B.; Fernández-Marcos, P.J.; Sánchez-Céspedes, 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]
  167. Mancarella, S.; Gigante, I.; Serino, G.; Pizzuto, E.; Dituri, F.; Valentini, M.F.; Wang, J.; Chen, X.; Armentano, R.; Calvisi, D.F.; et al. Crenigacestat blocking notch pathway reduces liver fibrosis in the surrounding ecosystem of intrahepatic CCA viaTGF-β inhibition. J. Exp. Clin. Cancer Res. 2022, 41, 331. [Google Scholar] [CrossRef]
  168. Abd El-Rhman, R.H.; El-Naga, R.N.; Gad, A.M.; Tadros, M.G.; Hassaneen, S.K. Dibenzazepine Attenuates Against Cisplatin-Induced Nephrotoxicity in Rats: Involvement of NOTCH Pathway. Front. Pharmacol. 2020, 11, 567852. [Google Scholar] [CrossRef]
  169. Alsemeh, A.; Abd El- Fatah, S.; Abdel Hamid, R.; Abbas, N.; Abdullah, D. Notch γ-Secretase Inhibitor Dibenzazepine Attenuates Cisplatin-induced Spleen Toxicity in Rats: Role of Notch Signaling Pathway. Zagazig Univ. Med. J. 2022, 28, 1242–1253. [Google Scholar] [CrossRef]
  170. Ahmed, L.A.; Abd El-Rhman, R.H.; Gad, A.M.; Hassaneen, S.K.; El-Yamany, M.F. Dibenzazepine combats acute liver injury in rats via amendments of Notch signaling and activation of autophagy. Naunyn-Schmiedebergs Arch. Pharmacol. 2021, 394, 337–348. [Google Scholar] [CrossRef]
  171. Wu, J.; Dong, X.; Li, W.; Zhao, L.; Zhou, L.; Sun, S.; Li, H. Dibenzazepine promotes cochlear supporting cell proliferation and hair cell regeneration in neonatal mice. Cell Prolif. 2020, 53, e12872. [Google Scholar] [CrossRef]
  172. Giordano, A.; Tommonaro, G. Curcumin and Cancer. Nutrients 2019, 11, 2376. [Google Scholar] [CrossRef]
  173. Hesari, A.R.; Azizian, M.; Sheikhi, A.; Nesaei, A.; Sanaei, S.; Mahinparvar, N.; Derakhshani, M.; Hedayt, P.; Ghasemi, F.; Mirzaei, H. Chemopreventive and therapeutic potential of curcumin in esophageal cancer: Current and future status. Int. J. Cancer 2019, 144, 1215–1226. [Google Scholar] [CrossRef]
  174. Liu, S.; Cao, Y.; Qu, M.; Zhang, Z.; Feng, L.; Ye, Z.; Xiao, M.; Hou, S.T.; Zheng, R.; Han, Z. Curcumin protects against stroke and increases levels of Notch intracellular domain. Neurol. Res. 2016, 38, 553–559. [Google Scholar] [CrossRef]
  175. Tandon, A.; Singh, S.J.; Gupta, M.; Singh, N.; Shankar, J.; Arjaria, N.; Goyal, S.; Chaturvedi, R.K. Notch pathway up-regulation via curcumin mitigates bisphenol-A (BPA) induced alterations in hippocampal oligodendrogenesis. J. Hazard. Mater. 2020, 392, 122052. [Google Scholar] [CrossRef] [PubMed]
  176. Zhdanovskaya, N.; Lazzari, S.; Caprioglio, D.; Firrincieli, M.; Maioli, C.; Pace, E.; Imperio, D.; Talora, C.; Bellavia, D.; Checquolo, S.; et al. Identification of a Novel Curcumin Derivative Influencing Notch Pathway and DNA Damage as a Potential Therapeutic Agent in T-ALL. Cancers 2022, 14, 5772. [Google Scholar] [CrossRef] [PubMed]
  177. Han, Z.; Zhang, J.; Zhang, K.; Zhao, Y. Curcumin inhibits cell viability, migration, and invasion of thymic carcinoma cells via downregulation of microRNA-27a. Phyther. Res. 2020, 34, 1629–1637. [Google Scholar] [CrossRef] [PubMed]
  178. Wu, G.Q.; Chai, K.Q.; Zhu, X.M.; Jiang, H.; Wang, X.; Xue, Q.; Zheng, A.H.; Zhou, H.Y.; Chen, Y.; Chen, X.C.; et al. Anti-cancer effects of curcumin on lung cancer through the inhibition of EZH2 and NOTCH1. Oncotarget 2016, 7, 26535–26550. [Google Scholar] [CrossRef]
  179. Sha, J.; Li, J.; Wang, W.; Pan, L.; Cheng, J.; Li, L.; Zhao, H.; Lin, W. Curcumin induces G0/G1 arrest and apoptosis in hormone independent prostate cancer DU-145 cells by down regulating Notch signaling. Biomed. Pharmacother. 2016, 84, 177–184. [Google Scholar] [CrossRef]
  180. Tang, Y.; Cao, Y. Curcumin Inhibits the Growth and Metastasis of Melanoma via miR-222-3p/SOX10/Notch Axis. Dis. Markers 2022, 2022, 3129781. [Google Scholar] [CrossRef]
  181. Naujokat, C.; McKee, D.L. The “Big Five” Phytochemicals Targeting Cancer Stem Cells: Curcumin, EGCG, Sulforaphane, Resveratrol and Genistein. Curr. Med. Chem. 2020, 28, 4321–4342. [Google Scholar] [CrossRef] [PubMed]
  182. Avila-Carrasco, L.; Majano, P.; Sánchez-Toméro, J.A.; Selgas, R.; López-Cabrera, M.; Aguilera, A.; González Mateo, G. Natural Plants Compounds as Modulators of Epithelial-to-Mesenchymal Transition. Front. Pharmacol. 2019, 10, 715. [Google Scholar] [CrossRef] [PubMed]
  183. Li, Y.; Zhang, T. Targeting cancer stem cells by curcumin and clinical applications. Cancer Lett. 2014, 346, 197–205. [Google Scholar] [CrossRef] [PubMed]
  184. Wang, Z.; Desmoulin, S.; Banerjee, S.; Kong, D.; Li, Y.; Deraniyagala, R.L.; Abbruzzese, J.; Sarkar, F.H. Synergistic effects of multiple natural products in pancreatic cancer cells. Life Sci. 2008, 83, 293–300. [Google Scholar] [CrossRef]
  185. Allen, J.L.; Hames, R.A.; Mastroianni, N.M.; Greenstein, A.E.; Weed, S.A. Evaluation of the matrix metalloproteinase 9 (MMP9) inhibitor Andecaliximab as an Anti-invasive therapeutic in Head and neck squamous cell carcinoma. Oral Oncol. 2022, 132, 106008. [Google Scholar] [CrossRef]
  186. Guo, Z.; Jin, X.; Jia, H. Inhibition of ADAM-17 more effectively down-regulates the Notch pathway than that of γ-secretase in renal carcinoma. J. Exp. Clin. Cancer Res. 2013, 32, 26. [Google Scholar] [CrossRef]
  187. Lu, H.Y.; Zu, Y.X.; Jiang, X.W.; Sun, X.T.; Liu, T.Y.; Li, R.L.; Wu, Q.; Zhang, Y.S.; Zhao, Q.C. Novel ADAM-17 inhibitor ZLDI-8 inhibits the proliferation and metastasis of chemo-resistant non-small-cell lung cancer by reversing Notch and epithelial mesenchymal transition in vitro and in vivo. Pharmacol. Res. 2019, 148, 104406. [Google Scholar] [CrossRef]
  188. Li, D.D.; Zhao, C.H.; Ding, H.W.; Wu, Q.; Ren, T.S.; Wang, J.; Chen, C.Q.; Zhao, Q.C. A novel inhibitor of ADAM17 sensitizes colorectal cancer cells to 5-Fluorouracil by reversing Notch and epithelial-mesenchymal transition in vitro and in vivo. Cell Prolif. 2018, 51, e12480. [Google Scholar] [CrossRef]
  189. Moss, M.L.; Minond, D. Recent Advances in ADAM17 Research: A Promising Target for Cancer and Inflammation. Mediat. Inflamm. 2017, 2017, 9673537. [Google Scholar] [CrossRef]
  190. Yang, L.; Bhattacharya, A.; Li, Y.; Sexton, S.; Ling, X.; Li, F.; Zhang, Y. Depleting receptor tyrosine kinases EGFR and HER2 overcomes resistance to EGFR inhibitors in colorectal cancer. J. Exp. Clin. Cancer Res. 2022, 41, 184. [Google Scholar] [CrossRef]
  191. Astudillo, L.; Da Silva, T.G.; Wang, Z.; Han, X.; Jin, K.; VanWye, J.; Zhu, X.; Weaver, K.; Oashi, T.; Lopes, P.E.M.; 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]
  192. Hao, N.; Yang, D.; Liu, T.; Liu, S.; Lu, X.; Chen, L. Laminin-integrin a6b4 interaction activates notch signaling to facilitate bladder cancer development. BMC Cancer 2022, 22, 558. [Google Scholar] [CrossRef] [PubMed]
  193. Kamga, P.T.; Bassi, G.; Cassaro, A.; Midolo, M.; Di Trapani, M.; Gatti, A.; Carusone, R.; Resci, F.; Perbellini, O.; Gottardi, M.; et al. Notch signalling drives bone marrow stromal cell-mediated chemoresistance in acute myeloid leukemia. Oncotarget 2016, 7, 21713–21727. [Google Scholar] [CrossRef]
  194. Kamga, P.T.; Dal Collo, G.; Midolo, M.; Adamo, A.; Delfino, P.; Mercuri, A.; Cesaro, S.; Mimiola, E.; Bonifacio, M.; Andreini, A.; et al. Inhibition of notch signaling enhances chemosensitivity in B-cell precursor acute lymphoblastic leukemia. Cancer Res. 2019, 79, 639–649. [Google Scholar] [CrossRef] [PubMed]
  195. Lehal, R.; Zaric, J.; Vigolo, M.; Urech, C.; Frismantas, V.; Zangger, N.; Cao, L.; Berger, A.; Chicote, I.; Loubéry, S.; et al. Pharmacological disruption of the Notch transcription factor complex. Proc. Natl. Acad. Sci. USA 2020, 117, 16292–16301. [Google Scholar] [CrossRef]
  196. Medinger, M.; Junker, T.; Heim, D.; Tzankov, A.; Jermann, P.M.; Bobadilla, M.; Vigolo, M.; Lehal, R.; Vogl, F.D.; Bauer, M.; et al. CB-103: A novel CSL-NICD inhibitor for the treatment of NOTCH-driven T-cell acute lymphoblastic leukemia: A case report of complete clinical response in a patient with relapsed and refractory T-ALL. eJHaem 2022, 3, 1009–1012. [Google Scholar] [CrossRef]
  197. Bui, T.O.; Angeli, E.; El Bouchtaoui, M.; Gapihan, G.; Dao, V.T.; Paris, J.; Leboeuf, C.; Soussan, M.; Villarese, P.; Ziol, M.; et al. Metastatic clear-cell renal cell carcinoma: A frequent NOTCH1 mutation predictive of response to anti-NOTCH1 CB-103 treatment. Exp. Hematol. Oncol. 2023, 12, 46. [Google Scholar] [CrossRef]
  198. Pinazza, M.; Ghisi, M.; Minuzzo, S.; Agnusdei, V.; Fossati, G.; Ciminale, V.; Pezzè, L.; Ciribilli, Y.; Pilotto, G.; Venturoli, C.; et al. Histone deacetylase 6 controls Notch3 trafficking and degradation in T-cell acute lymphoblastic leukemia cells. Oncogene 2018, 37, 3839–3851. [Google Scholar] [CrossRef]
  199. Hurtado, C.; Safarova, A.; Smith, M.; Chung, R.; Bruyneel, A.A.N.; Gomez-Galeno, J.; Oswald, F.; Larson, C.J.; Cashman, J.R.; Ruiz-Lozano, P.; et al. Disruption of NOTCH signaling by a small molecule inhibitor of the transcription factor RBPJ. Sci. Rep. 2019, 9, 10811. [Google Scholar] [CrossRef]
  200. Yuan, Z.; VanderWielen, B.D.; Giaimo, B.D.; Pan, L.; Collins, C.E.; Turkiewicz, A.; Hein, K.; Oswald, F.; Borggrefe, T.; Kovall, R.A. Structural and Functional Studies of the RBPJ-SHARP Complex Reveal a Conserved Corepressor Binding Site. Cell Rep. 2019, 26, 845–854. [Google Scholar] [CrossRef]
  201. Oswald, F.; Rodriguez, P.; Giaimo, B.D.; Antonello, Z.A.; Mira, L.; Mittler, G.; Thiel, V.N.; Collins, K.J.; Tabaja, N.; Cizelsky, W.; et al. A phospho-dependent mechanism involving NCoR and KMT2D controls a permissive chromatin state at Notch target genes. Nucleic Acids Res. 2016, 44, 4703–4720. [Google Scholar] [CrossRef]
  202. Feng, S.; Yang, Y.; Lv, J.; Sun, L.; Liu, M. Valproic acid exhibits different cell growth arrest effect in three HPV-positive/negative cervical cancer cells and possibly via inducing Notch1 cleavage and E6 downregulation. Int. J. Oncol. 2016, 49, 422–430. [Google Scholar] [CrossRef] [PubMed]
  203. Sun, L.; Qian, Q.; Sun, G.; Mackey, L.V.; Fuselier, J.A.; Coy, D.H.; Yu, C.Y. Valproic acid induces NET cell growth arrest and enhances tumor suppression of the receptor-targeted peptide-drug conjugate via activating somatostatin receptor type II. J. Drug Target. 2016, 24, 169–177. [Google Scholar] [CrossRef] [PubMed]
  204. Aljedai, A.; Buckle, A.-M.; Hiwarkar, P.; Syed, F. Potential Role of Notch Signalling in CD34+ Chronic Myeloid Leukaemia Cells: Cross-Talk between Notch and BCR-ABL. PLoS ONE 2015, 10, e0123016. [Google Scholar] [CrossRef] [PubMed]
  205. Yang, X.; Liu, J.; Liang, Q.; Sun, G. Valproic acid reverses sorafenib resistance through inhibiting activated Notch/Akt signaling pathway in hepatocellular carcinoma. Fundam. Clin. Pharmacol. 2021, 35, 690–699. [Google Scholar] [CrossRef]
  206. Singh, D.; Gupta, S.; Singh, I.; Morsy, M.A.; Nair, A.B.; Ahmed, A.S.F. Hidden pharmacological activities of valproic acid: A new insight. Biomed. Pharmacother. 2021, 142, 112021. [Google Scholar] [CrossRef]
  207. LaFoya, B.; Munroe, J.A.; Albig, A.R. A comparison of resveratrol and other polyphenolic compounds on Notch activation and endothelial cell activity. PLoS ONE 2019, 14, e0210607. [Google Scholar] [CrossRef]
  208. Liu, X.; Zhang, Q.; Li, Y.; Zhao, D.X.; Gu, R. Dual mechanism of action of resveratrol in notch signaling pathway activation in osteosarcoma. Trop. J. Pharm. Res. 2016, 15, 101–106. [Google Scholar] [CrossRef]
  209. Pinchot, S.N.; Jaskula-Sztul, R.; Ning, L.; Peters, N.R.; Cook, M.R.; Kunnimalaiyaan, M.; Chen, H. Identification and validation of Notch pathway activating compounds through a novel high-throughput screening method. Cancer 2011, 117, 1386–1398. [Google Scholar] [CrossRef]
  210. Truong, M.; Cook, M.R.; Pinchot, S.N.; Kunnimalaiyaan, M.; Chen, H. Resveratrol induces Notch2-mediated apoptosis and suppression of neuroendocrine markers in medullary thyroid cancer. Ann. Surg. Oncol. 2011, 18, 1506–1511. [Google Scholar] [CrossRef]
  211. Yu, X.M.; Jaskula-Sztul, R.; Ahmed, K.; Harrison, A.D.; Kunnimalaiyaan, M.; Chen, H. Resveratrol induces differentiation markers expression in anaplastic thyroid carcinoma via activation of Notch1 signaling and suppresses cell growth. Mol. Cancer Ther. 2013, 12, 1276–1287. [Google Scholar] [CrossRef] [PubMed]
  212. Dong, J.; Yang, W.; Han, J.; Cheng, R.; Li, L. Effects of Notch signaling components from breast cancer cells treated in culture with resveratrol. Res. Vet. Sci. 2020, 132, 369–378. [Google Scholar] [CrossRef] [PubMed]
  213. Zhong, L.X.; Li, H.; Wu, M.L.; Liu, X.Y.; Zhong, M.J.; Chen, X.Y.; Liu, J.; Zhang, Y. Inhibition of STAT3 signaling as critical molecular event in resveratrol-suppressed ovarian cancer cells. J. Ovarian Res. 2015, 8, 25. [Google Scholar] [CrossRef] [PubMed]
  214. Kim, T.H.; Park, J.H.; Woo, J.S. Resveratrol induces cell death through ROS-dependent downregulation of Notch1/PTEN/Akt signaling in ovarian cancer cells. Mol. Med. Rep. 2019, 19, 3353–3360. [Google Scholar] [CrossRef]
  215. Venturelli, S.; Berger, A.; Böcker, A.; Busch, C.; Weiland, T.; Noor, S.; Leischner, C.; Schleicher, S.; Mayer, M.; Weiss, T.S.; et al. Resveratrol as a Pan-HDAC Inhibitor Alters the Acetylation Status of Jistone Proteins in Human-Derived Hepatoblastoma Cells. PLoS ONE 2013, 8, e73097. [Google Scholar] [CrossRef]
  216. Ren, B.; Kwah, M.X.Y.; Liu, C.; Ma, Z.; Shanmugam, M.K.; Ding, L.; Xiang, X.; Ho, P.C.L.; Wang, L.; Ong, P.S.; et al. Resveratrol for cancer therapy: Challenges and future perspectives. Cancer Lett. 2021, 515, 63–72. [Google Scholar] [CrossRef]
  217. Koyama, D.; Kikuchi, J.; Hiraoka, N.; Wada, T.; Kurosawa, H.; Chiba, S.; Furukawa, Y. Proteasome inhibitors exert cytotoxicity and increase chemosensitivity via transcriptional repression of Notch1 in T-cell acute lymphoblastic leukemia. Leukemia 2014, 28, 1216–1226. [Google Scholar] [CrossRef]
  218. Yu, P.; Petrus, M.N.; Ju, W.; Zhang, M.; Conlon, K.C.; Nakagawa, M.; Maeda, M.; Bamford, R.N.; Waldmann, T.A. Augmented efficacy with the combination of blockade of the Notch-1 pathway, bortezomib and romidepsin in a murine MT-1 adult T-cell leukemia model. Leukemia 2015, 29, 556–566. [Google Scholar] [CrossRef]
  219. Thounaojam, M.C.; Dudimah, D.F.; Pellom, S.T.; Uzhachenko, R.V.; Carbone, D.P.; Dikov, M.M.; Shanker, A. Bortezomib enhances expression of effector molecules in antitumor CD8+T lymphocytes by promoting Notch-nuclear factor-κB crosstalk. Oncotarget 2015, 6, 32439–32455. [Google Scholar] [CrossRef]
  220. Clementz, A.G.; Osipo, C. Notch versus the proteasome: What is the target of γ-secretase inhibitor-I? Breast Cancer Res. 2009, 11, 110. [Google Scholar] [CrossRef]
  221. Wu, M.H.; Lu, R.Y.; Yu, S.J.; Tsai, Y.Z.; Lin, Y.C.; Bai, Z.Y.; Liao, R.Y.; Hsu, Y.C.; Chen, C.C.; Cai, B.H. PTC124 Rescues Nonsense Mutation of Two Tumor Suppressor Genes NOTCH1 and FAT1 to Repress HNSCC Cell Proliferation. Biomedicines 2022, 10, 2948. [Google Scholar] [CrossRef] [PubMed]
  222. Welch, E.M.; Barton, E.R.; Zhuo, J.; Tomizawa, Y.; Friesen, W.J.; Trifillis, P.; Paushkin, S.; Patel, M.; Trotta, C.R.; Hwang, S.; et al. PTC124 targets genetic disorders caused by nonsense mutations. Nature 2007, 447, 87–91. [Google Scholar] [CrossRef] [PubMed]
  223. Siddiqui, N.; Sonenberg, N. Proposing a mechanism of action for ataluren. Proc. Natl. Acad. Sci. USA 2016, 113, 12353–12355. [Google Scholar] [CrossRef] [PubMed]
  224. Tutone, M.; Pibiri, I.; Lentini, L.; Pace, A.; Almerico, A.M. Deciphering the Nonsense Readthrough Mechanism of Action of Ataluren: An in Silico Compared Study. ACS Med. Chem. Lett. 2019, 10, 522–527. [Google Scholar] [CrossRef]
  225. Huang, S.; Bhattacharya, A.; Ghelfi, M.D.; Li, H.; Fritsch, C.; Chenoweth, D.M.; Goldman, Y.E.; Cooperman, B.S. Ataluren binds to multiple protein synthesis apparatus sites and competitively inhibits release factor-dependent termination. Nat. Commun. 2022, 13, 2413. [Google Scholar] [CrossRef]
  226. Du, M.; Liu, X.; Welch, E.M.; Hirawat, S.; Peltz, S.W.; Bedwell, D.M. PTC124 is an orally bioavailable compound that promotes suppression of the human CFTR-G542X nonsense allele in a CF mouse model. Proc. Natl. Acad. Sci. USA 2008, 105, 2064–2069. [Google Scholar] [CrossRef]
  227. Rui, M.; Cai, M.; Zhou, Y.; Zhang, W.; Gao, L.; Mi, K.; Ji, W.; Wang, D.; Feng, C. Identification of Potential RBPJ-Specific Inhibitors for Blocking Notch Signaling in Breast Cancer Using a Drug Repurposing Strategy. Pharmaceuticals 2022, 15, 556. [Google Scholar] [CrossRef]
  228. Sen, P.; Kandasamy, T.; Ghosh, S.S. Multi-targeting TACE/ADAM17 and gamma-secretase of notch signalling pathway in TNBC via drug repurposing approach using Lomitapide. Cell. Signal. 2023, 102, 110529. [Google Scholar] [CrossRef]
  229. 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]
  230. Ferrarotto, R.; Eckhardt, G.; Patnaik, A.; LoRusso, P.; Faoro, L.; Heymach, J.V.; Kapoun, A.M.; Xu, L.; Munster, P. A phase I dose-escalation and dose-expansion study of brontictuzumab in subjects with selected solid tumors. Ann. Oncol. 2018, 29, 1561–1568. [Google Scholar] [CrossRef]
  231. Xie, M.; Wei, S.; Wu, X.; Li, X.; You, Y.; He, C. Alterations of Notch pathway in patients with adenoid cystic carcinoma of the trachea and its impact on survival. Lung Cancer 2018, 121, 41–47. [Google Scholar] [CrossRef] [PubMed]
  232. Ferrarotto, R.; Mitani, Y.; Diao, L.; Guijarro, I.; Wang, J.; Zweidler-McKay, P.; Bell, D.; William, W.N.; Glisson, B.S.; Wick, M.J.; et al. Activating NOTCH1 mutations define a distinct subgroup of patients with adenoid cystic carcinoma who have poor prognosis, propensity to bone and liver metastasis, and potential responsiveness to Notch1 inhibitors. J. Clin. Oncol. 2017, 35, 352–360. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Canonical and non-canonical Notch signaling pathway. A schematic figure of the structure of the receptors (NOTCH1-4) and their ligands (DLL-1/3/4 and JAG1/2). After synthesis, the NOTCH receptors mature in a stepwise process and are simultaneously transported through the endoplasmic reticulum and Golgi structures to the plasma membrane. Receptor–ligand interactions generate a mechanical pulling force that allows ADAM proteases and the γ-secretase complex to cleave the receptor (at the S2 and S3 cleave site, respectively), followed by the release of the intracellular domain (NICD) to the cytoplasm of the signal-receiving cell. Simultaneously, the extracellular part of the NOTCH (NECD) and the ligand are internalized by the signal-sending cell. Intracellular release of NICD can also occur without ligand activation via NOTCH internalization and γ-secretase activity. The NICD is translocated into the nucleus where it participates in the formation of the Notch ternary complex (NTC), activating the expression of target genes. In non-canonical Notch signaling, the NICD interacts with and modulates other signaling pathways to express target genes without the mediation of the NTC.
Figure 1. Canonical and non-canonical Notch signaling pathway. A schematic figure of the structure of the receptors (NOTCH1-4) and their ligands (DLL-1/3/4 and JAG1/2). After synthesis, the NOTCH receptors mature in a stepwise process and are simultaneously transported through the endoplasmic reticulum and Golgi structures to the plasma membrane. Receptor–ligand interactions generate a mechanical pulling force that allows ADAM proteases and the γ-secretase complex to cleave the receptor (at the S2 and S3 cleave site, respectively), followed by the release of the intracellular domain (NICD) to the cytoplasm of the signal-receiving cell. Simultaneously, the extracellular part of the NOTCH (NECD) and the ligand are internalized by the signal-sending cell. Intracellular release of NICD can also occur without ligand activation via NOTCH internalization and γ-secretase activity. The NICD is translocated into the nucleus where it participates in the formation of the Notch ternary complex (NTC), activating the expression of target genes. In non-canonical Notch signaling, the NICD interacts with and modulates other signaling pathways to express target genes without the mediation of the NTC.
Cancers 15 04563 g001
Figure 2. Modulation of Notch signaling by interfering with membrane trafficking of NOTCH receptors. Inhibition of the membrane ion transporters SLC39A7 (NVS-ZP7-4) and SERCA (Thapsigargin, CAD204520, Casearin J) disrupts Zn2+ and Ca2+ ion efflux between endoplasmic reticulum/Golgi structures and the cytoplasm, resulting in impaired protein folding, arrested maturation, and a block of intracellular NOTCH protein transport. Attenuation of receptor exposure to the membrane can also be achieved by compounds blocking NOTCH secretion prior to the exit from the endoplasmic reticulum (FLI-06), or leading to the accumulation of NOTCH receptors in autophagosomes (Chloroquine, more details in the text). Red: attenuation of Notch signaling (NVS-ZP7-4, Thapsigargin, CAD204520, Casearin J, Bepridil, FLI-06, Chloroquine).
Figure 2. Modulation of Notch signaling by interfering with membrane trafficking of NOTCH receptors. Inhibition of the membrane ion transporters SLC39A7 (NVS-ZP7-4) and SERCA (Thapsigargin, CAD204520, Casearin J) disrupts Zn2+ and Ca2+ ion efflux between endoplasmic reticulum/Golgi structures and the cytoplasm, resulting in impaired protein folding, arrested maturation, and a block of intracellular NOTCH protein transport. Attenuation of receptor exposure to the membrane can also be achieved by compounds blocking NOTCH secretion prior to the exit from the endoplasmic reticulum (FLI-06), or leading to the accumulation of NOTCH receptors in autophagosomes (Chloroquine, more details in the text). Red: attenuation of Notch signaling (NVS-ZP7-4, Thapsigargin, CAD204520, Casearin J, Bepridil, FLI-06, Chloroquine).
Cancers 15 04563 g002
Figure 3. Modulation of NICD release. ADAM protease inhibitors (Lomitapide, Aderbasib, Marimastat, ZLDI-8) block NOTCH S2 cleavage and, consequently, further activation of the NOTCH receptor. The use of γ-secretase inhibitors (γSI; GSI) prevents S3 cleavage and subsequent NICD release. The NICD can be translocated to the nucleus where it is involved in canonical Notch signaling or interacts with other signaling pathways as a component of non-canonical Notch signaling. These events are influenced by proteasome activity (inhibited by Bortezomib) and a number of broad-spectrum substances collectively named as Pan-Assay INterference compoundS (PAINS) such as Resveratrol and Curcumin. Red: attenuation of Notch signaling; Blue: pleiotropic effect on Notch signaling.
Figure 3. Modulation of NICD release. ADAM protease inhibitors (Lomitapide, Aderbasib, Marimastat, ZLDI-8) block NOTCH S2 cleavage and, consequently, further activation of the NOTCH receptor. The use of γ-secretase inhibitors (γSI; GSI) prevents S3 cleavage and subsequent NICD release. The NICD can be translocated to the nucleus where it is involved in canonical Notch signaling or interacts with other signaling pathways as a component of non-canonical Notch signaling. These events are influenced by proteasome activity (inhibited by Bortezomib) and a number of broad-spectrum substances collectively named as Pan-Assay INterference compoundS (PAINS) such as Resveratrol and Curcumin. Red: attenuation of Notch signaling; Blue: pleiotropic effect on Notch signaling.
Cancers 15 04563 g003
Figure 4. Modulation of Notch signaling by targeting the NICD-dependent transcription complex. Attenuation of Notch signaling can be achieved by blocking ternary complex formation by inhibiting RBP-J activity (Fidaxamicin) or preventing the binding of MAML to NICD/RPB-J (IMR-1A, SAHM1) or NICD to RBP-J (CB-103). RIN-1 blocks the binding of the SHARP co-repressor to RBP-j, preventing the function of the repressor complex. HDAC inhibitors (Givinostat, Trichostatin, Tubacin, Valproic acid) block the ability of HDACs to deacetylate histones, contributing to the maintenance of an open DNA structure and expression of genes. Finally, some compounds, such as PTC-124, can re-express functional NOTCH1 and induce Notch-regulated gene expression levels. Red: attenuation of Notch signaling (IMR-1, SAHM1, CB-103, Tubacin), Blue: pleiotropic effect on Notch signaling (RIN-1, VPA), Green: activation of Notch signaling.
Figure 4. Modulation of Notch signaling by targeting the NICD-dependent transcription complex. Attenuation of Notch signaling can be achieved by blocking ternary complex formation by inhibiting RBP-J activity (Fidaxamicin) or preventing the binding of MAML to NICD/RPB-J (IMR-1A, SAHM1) or NICD to RBP-J (CB-103). RIN-1 blocks the binding of the SHARP co-repressor to RBP-j, preventing the function of the repressor complex. HDAC inhibitors (Givinostat, Trichostatin, Tubacin, Valproic acid) block the ability of HDACs to deacetylate histones, contributing to the maintenance of an open DNA structure and expression of genes. Finally, some compounds, such as PTC-124, can re-express functional NOTCH1 and induce Notch-regulated gene expression levels. Red: attenuation of Notch signaling (IMR-1, SAHM1, CB-103, Tubacin), Blue: pleiotropic effect on Notch signaling (RIN-1, VPA), Green: activation of Notch signaling.
Cancers 15 04563 g004
Table 1. Clinical trials of small molecular compounds targeting Notch signaling.
Table 1. Clinical trials of small molecular compounds targeting Notch signaling.
Study TitleNCT NumberStatusConditionsInterventions
A Study Of AL101 in Patients with Adenoid Cystic Carcinoma (ACC) Bearing Activating Notch MutationsNCT03691207Unknown statusAdenoid Cystic CarcinomaAL101
AL101 Before Surgery for the Treatment of Notch-Activated Adenoid Cystic CancerNCT04973683RecruitingAdenoid Cystic CarcinomaAL101
A Study of AL101 Monotherapy in Patients with Notch-Activated Triple-Negative Breast CancerNCT04461600Active, not recruitingTriple-Negative Breast CancerAL101
A Study of AL102 in Patients with Progressing Desmoid TumorsNCT04871282RecruitingDesmoid TumorAL102
Single-arm Study with Bimiralisib in Patients With HNSCC Harboring NOTCH1 Loss of Function MutationsNCT03740100TerminatedHNSCCBimiralisib (PI3Kinase inhibitor)
Study to Evaluate the Safety and Tolerability of Weekly Intravenous (IV) Doses of BMS-906024 in Subjects with Acute T-cell Lymphoblastic Leukemia or T-cell Lymphoblastic LymphomaNCT01363817CompletedAcute T-cel Lymphoblastic Leukemia, Precursor T-Cell LeukemiaBMS-906024 (γ-secretase inhibitor)
Study to Evaluate the Safety and Tolerability of IV Doses of BMS-906024 in Subjects with Advanced or Metastatic Solid TumorsNCT01292655CompletedCancerBMS-906024
Study to Evaluate Safety and Tolerability of BMS-906024 in Combination with ChemotherapyNCT01653470CompletedSolid cancersBMS-906024, Paclitaxel, 5-Fluorouracil (5FU), Carboplatin
Phase I Ascending Multiple-Dose Study of BMS-986115 in Subjects with Advanced Solid TumorsNCT01986218TerminatedVarious Advanced CancersBMS-986115 (γ-secretase inhibitor)
Compassionate Use of Brontictuzumab for Adenoid Cystic Carcinoma (ACC)NCT02662608CompletedAdenoid Cystic CarcinomaBrontictuzumab (monoclonal antibody that targets Notch1)
CB-103 Plus NSAI in Luminal Advanced Breast CancerNCT04714619Active, not recruitingAdvanced Breast CancerCB-103, NSAIDs
Study of CB-103 in Adult Patients with Advanced or Metastatic Solid Tumors and Hematological MalignanciesNCT03422679TerminatedBreast Cancer, Colorectal, ACC, Osteosarcoma, HCCCB-103
A Phase 1/2 Study CB-103 with or without Venetoclax in Patients with NOTCH ACCNCT05774899RecruitingAdenoid Cystic CarcinomaCB-103, Venetoclax, Lenvatinib
Study in Patients with Advanced Cancers Associated with Expression of DLL3NCT04471727RecruitingSmall-cell Lung CancerHPN328 (a DLL3-targeting T-cell engager), Atezolizumab
A Study of LY3039478 in Japanese Participants with Advanced Solid TumorsNCT02836600Active, not recruitingAdvanced Solid TumorCrenigacestat/LY3039478
A Study of LY3039478 in Participants with Advanced or Metastatic Solid TumorsNCT02784795CompletedSolid Tumors: Breast Cancer, Colon CancerCrenigacestat/LY3039478, Taladegib, Abemaciclib, Cisplatin, Gemcitabine, Carboplatin
Notch Inhibitor in Advanced CancerNCT01158404CompletedAdvanced CancerLY900009 (γ-secretase inhibitor)
A Phase 1 Study to Evaluate the Safety, Tolerability, and Pharmacokinetics of MEDI0639 in Advanced Solid TumorsNCT01577745CompletedSolid TumorsMEDI0639 (DLL4-targeting antibody)
A Notch Signaling Pathway Inhibitor for Patients with Advanced Breast Cancer (0752-014)NCT00106145CompletedAdvanced Breast Cancer, other Solid TumorsMK-0752
A Notch Signaling Pathway Inhibitor for Patients with T-cell Acute Lymphoblastic Leukemia/Lymphoma (ALL) (0752-013)NCT00100152TerminatedMyelogenous Leukemia, Chronic Lymphocytic Leukemia, Lymphoblastic Acute T-cell LeukemiaMK-0752
Phase I/II Study of MK-0752 Followed by Docetaxel in Advanced or Metastatic Breast CancerNCT00645333CompletedMetastatic Breast CancerMK-0752, Docetaxel, Pegfilgrastim
MK-0752 and Gemcitabine Hydrochloride in Treating Patients with Stage III and IV Pancreatic Cancer that Cannot be Removed by SurgeryNCT01098344CompletedPancreatic CancerMK-0752, gemcitabine
Nirogacestat for Adults with Desmoid Tumor/Aggressive Fibromatosis (DT/AF)NCT03785964Active, not recruitingDesmoid Tumor, Aggressive FibromatosisNirogacestat
Individual Patient Compassionate Use of NirogacestatNCT05041036AvailableDesmoid Tumors, NOTCH Gene Mutation Positive TumorsNirogacestat
γ-secretase inhibitor PF-03084014 in Treating Patients with AIDS-Associated Kaposi SarcomaNCT02137564WithdrawnAIDS-related Kaposi SarcomaPF-03084014
Biomarker Research Study for PF-03084014 in Chemo-resistant Triple-negative Breast CancerNCT02338531WithdrawnBreast CancerPF-03084014
A Study Evaluating The PF-03084014 in Combination with Docetaxel in Patients with Advanced Breast CancerNCT01876251TerminatedBreast Cancer MetastaticPF-03084014
Phase II Trial of the γ-Secretase Inhibitor PF-03084014 in Adults With Desmoid Tumors/Aggressive FibromatosisNCT01981551Active, not recruitingDesmoid Tumors, Aggressive FibromatosisPF-03084014
A Trial in Patients with Advanced Cancer and LeukemiaNCT00878189CompletedAdvanced Cancer And LeukemiaPF-03084014
A Study Evaluating PF-03084014 in Patients with Advanced Breast Cancer with Or without Notch AlterationsNCT02299635TerminatedTriple-Negative Breast NeoplasmsPF-03084014
γ-secretase inhibitor RO4929097 in Previously Treated Metastatic Pancreas CancerNCT01232829CompletedAdenocarcinoma of the Pancreas, Recurrent Pancreatic CancerRO4929097
RO4929097 Before Surgery in Treating Patients with Pancreatic CancerNCT01192763TerminatedAdenocarcinoma of the Pancreas, Stage IA, IB Pancreatic CancerRO4929097
Vismodegib and γ-secretase/Notch Signaling Pathway Inhibitor RO4929097 in Treating Patients with Advanced or Metastatic SarcomaNCT01154452CompletedAdult Alveolar Soft Part Sarcoma, Angiosarcoma, Desmoplastic Small Round Cell TumorRO4929097
RO4929097 in Treating Patients With Recurrent Invasive GliomasNCT01269411TerminatedAdult Anaplastic Oligodendroglioma, Brain Stem Glioma, Giant Cell GlioblastomaRO4929097
γ-secretase/Notch Signaling Pathway Inhibitor RO4929097 in Treating Patients with Recurrent or Progressive GlioblastomaNCT01122901TerminatedAdult Giant Cell Glioblastoma, GliosarcomaRO4929097
γ-secretase inhibitor RO4929097 in Treating Young Patients with Relapsed or Refractory Solid Tumors, CNS Tumors, Lymphoma, or T-Cell LeukemiaNCT01088763TerminatedChildhood Atypical Teratoid/Rhabdoid Tumor, Choriocarcinoma, GerminomaRO4929097
A Study of RO4929097 in Patients with Advanced Renal Cell Carcinoma that Have Failed Vascular Endothelial Growth Factor (VEGF)/Vascular Endothelial Growth Factor Receptor (VEGFR) TherapyNCT01141569CompletedClear Cell Renal Cell Carcinoma, recurrent, and Stage IV Renal Cell CancerRO4929097
γ-secretase/Notch Signaling Pathway Inhibitor RO4929097 in Treating Patients with Advanced, Metastatic, or Recurrent Triple-Negative Invasive Breast CancerNCT01151449TerminatedTriple-Negative Breast CancerRO4929097
γ-secretase/Notch Signaling Pathway Inhibitor RO4929097 in Treating Patients with Stage IV MelanomaNCT01120275TerminatedMalignant MelanomaRO4929097
RO4929097 in Treating Patients with Metastatic Colorectal CancerNCT01116687CompletedRecurrent Colon, Rectal Cancer, Stage IV Colon CancerRO4929097
RO4929097 in Treating Patients with Recurrent and/or Metastatic Epithelial Ovarian Cancer, Fallopian Tube Cancer, or Primary Peritoneal CancerNCT01175343CompletedFallopian Tube Carcinoma, Ovarian Carcinoma, Primary Peritoneal CarcinomaRO4929097
RO4929097 in Treating Patients with Advanced Non-Small Cell Lung Cancer who have Recently Completed Treatment with Front-Line ChemotherapyNCT01193868TerminatedRecurrent Non-Small Cell Lung Cancer, Stage IIIB, Stage IVRO4929097
RO4929097 in Treating Patients with Stage IIIB, Stage IIIC, or Stage IV Melanoma that Can be Removed by SurgeryNCT01216787WithdrawnMelanoma Stage IIIB, IIIC, IVRO4929097
γ-secretase inhibitor RO4929097 in Treating Patients with Metastatic or Unresectable Solid MalignanciesNCT01096355CompletedUnspecified Adult Solid TumorRO4929097
RO4929097 After Autologous Stem Cell Transplant in Treating Patients with Multiple MyelomaNCT01251172WithdrawnPlasma Cell Myeloma Stage I, II, IIIRO4929097, Autologous Hematopoietic Stem Cell Transplantation
RO4929097 and Bevacizumab in Treating Patients with Progressive or Recurrent Malignant GliomaNCT01189240TerminatedAdult Anaplastic Astrocytoma, Oligodendroglioma, Giant Cell GlioblastomaRO4929097, bevacizumab
Combination Chemotherapy and Bevacizumab with or without RO4929097 in Treating Patients with Metastatic Colorectal CancerNCT01270438WithdrawnAdenocarcinoma of the Colon, RectumRO4929097, bevacizumab, FOLFOX regimen,
Bicalutamide and RO4929097 in Treating Patients with Previously Treated Prostate CancerNCT01200810TerminatedAdenocarcinoma of the ProstateRO4929097, bicalutamide
RO4929097 and Capecitabine in Treating Patients with Refractory Solid TumorsNCT01158274CompletedAdult Grade III Lymphomatoid Granulomatosis, Adult Nasal Type Extranodal NK/T-cell Lymphoma, AIDS-related Diffuse Large Cell LymphomaRO4929097, capecitabine
γ-secretase/Notch Signaling Pathway Inhibitor RO4929097, Paclitaxel, and Carboplatin Before Surgery in Treating Patients with Stage II or Stage III Triple-Negative Breast CancerNCT01238133TerminatedTriple-Negative Breast CancerRO4929097, Carboplatin
Phase I Study of Cetuximab with RO4929097 in Metastatic Colorectal CancerNCT01198535TerminatedColon Mucinous Adenocarcinoma, Colon Signet Ring Cell Adenocarcinoma, Rectal Mucinous AdenocarcinomaRO4929097, Cetuximab
γ-secretase/Notch Signaling Pathway Inhibitor RO4929097 in Combination with Cisplatin, Vinblastine, and Temozolomide in Treating Patients with Recurrent or Metastatic MelanomaNCT01196416CompletedRecurrent Melanoma, Stage IV Skin MelanomaRO4929097, Cisplatin
RO4929097 in Children with Relapsed/Refractory Solid or CNS Tumors, Lymphoma, or T-Cell LeukemiaNCT01236586WithdrawnLymphomaRO4929097, Dexamethasone
RO4929097 and Erlotinib Hydrochloride in Treating Patients with Stage IV or Recurrent Non-Small Cell Lung CancerNCT01193881TerminatedRecurrent Non-Small Cell Lung Carcinoma, Stage IV Non-Small Cell Lung CancerRO4929097, Erlotinib Hydrochloride
γ-secretase Inhibitor RO4929097 and Gemcitabine Hydrochloride in Treating Patients with Advanced Solid TumorsNCT01145456CompletedAdenocarcinoma of the Pancreas, Recurrent Pancreatic Cancer, Stage IIIRO4929097, gemcitabine
γ-secretase/Notch Signaling Pathway Inhibitor RO4929097 in Treating Patients with Advanced Solid TumorsNCT01218620CompletedAdult Solid NeoplasmRO4929097, Ketoconazole, Rifampicin
RO4929097 and Letrozole in Treating Post-Menopausal Women with Hormone-Receptor-Positive Stage II or Stage III Breast CancerNCT01208441TerminatedEstrogen-Receptor-positive and Progesterone-Receptor-positive Breast cancer, HER2-negative Breast CancerRO4929097, letrozole
γ-secretase/Notch Signaling Pathway Inhibitor RO4929097 and Temsirolimus in Treating Patients with Advanced Solid TumorsNCT01198184CompletedEndometrial Papillary Serous Carcinoma, Recurrent Endometrial Carcinoma, Recurrent Renal Cell CancerRO4929097, temsirolimus
RO4929097 and Vismodegib in Treating Patients with Breast Cancer that is Metastatic or Cannot be Removed by SurgeryNCT01071564TerminatedTriple-Negative Breast CancerRO4929097, Vismodegib
RO4929097 and Whole-Brain Radiation Therapy or Stereotactic Radiosurgery in Treating Patients with Brain Metastases From Breast CancerNCT01217411TerminatedTriple-Negative Breast CancerRO4929097, Whole-brain radiation therapy (WBRT), Stereotactic radiosurgery (SRS)
RO4929097, Temozolomide, and Radiation Therapy in Treating Patients with Newly Diagnosed Malignant GliomaNCT01119599CompletedAcoustic Schwannoma, Anaplastic Astrocytoma, Anaplastic MeningiomaRO4929097, 3-Dimensional Conformal and Intensity-Modulated Radiation Therapy
A Phase 1b/2 Study of OMP-59R5 (Tarextumab) in Combination with Nab-Paclitaxel and Gemcitabine in Subjects with Previously Untreated Stage IV Pancreatic CancerNCT01647828CompletedPancreatic Cancer Stage IVTarextumab (OMP-59R5, a NOCTH2/3 targeting antibody), Gemcitabine
A Study of Tarlatamab in Participants with Neuroendocrine Prostate CancerNCT04702737Active, not recruitingNeuroendocrine Prostate CancerTarlatamab (a DLL3-targeting T-cell engager)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Czerwonka, A.; Kałafut, J.; Nees, M. Modulation of Notch Signaling by Small-Molecular Compounds and Its Potential in Anticancer Studies. Cancers 2023, 15, 4563. https://doi.org/10.3390/cancers15184563

AMA Style

Czerwonka A, Kałafut J, Nees M. Modulation of Notch Signaling by Small-Molecular Compounds and Its Potential in Anticancer Studies. Cancers. 2023; 15(18):4563. https://doi.org/10.3390/cancers15184563

Chicago/Turabian Style

Czerwonka, Arkadiusz, Joanna Kałafut, and Matthias Nees. 2023. "Modulation of Notch Signaling by Small-Molecular Compounds and Its Potential in Anticancer Studies" Cancers 15, no. 18: 4563. https://doi.org/10.3390/cancers15184563

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop