Next Article in Journal
Mutual Interaction of Clinical Factors and Specific microRNAs to Predict Mild Cognitive Impairment in Patients Receiving Hemodialysis
Next Article in Special Issue
Notch Signaling Function in the Angiocrine Regulation of Tumor Development
Previous Article in Journal
Mechanistic Modeling of the Relative Biological Effectiveness of Boron Neutron Capture Therapy
Previous Article in Special Issue
Targeting Notch Trafficking and Processing in Cancers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oncogenic and Tumor-Suppressive Functions of NOTCH Signaling in Glioma

Department of Biomedicine, University of Basel, Mattenstrasse 28, 4058 Basel, Switzerland
*
Author to whom correspondence should be addressed.
Cells 2020, 9(10), 2304; https://doi.org/10.3390/cells9102304
Submission received: 22 September 2020 / Revised: 13 October 2020 / Accepted: 14 October 2020 / Published: 15 October 2020
(This article belongs to the Special Issue The Roles of Notch Signaling in Cancers)

Abstract

:
Although the role of NOTCH signaling has been extensively studied in health and disease, many questions still remain unresolved. Being crucial for tissue homeostasis, NOTCH signaling is also implicated in multiple cancers by either promoting or suppressing tumor development. In this review we illustrate the context-dependent role of NOTCH signaling during tumorigenesis with a particular focus on gliomas, the most frequent and aggressive brain tumors in adults. For a long time, NOTCH has been considered an oncogene in glioma mainly by virtue of its neural stem cell-promoting activity. However, the recent identification of NOTCH-inactivating mutations in some glioma patients has challenged this notion, prompting a re-examination of the function of NOTCH in brain tumor subtypes. We discuss recent findings that might help to reconcile the controversial role of NOTCH signaling in this disease, and pose outstanding questions that still remain to be addressed.

1. Context-Dependent Roles of NOTCH Signaling in Cancer

The NOTCH signaling pathway transduces short-range signals between adjacent cells and therefore is highly dependent on niche architecture. In mammals, signal activation is induced by direct interaction of one of the four NOTCH receptors (NOTCH1–4) with one of five canonical ligands (Jagged1, Jagged2, Delta-like 1 (DLL1), DLL3, DLL4), followed by two sequential proteolytic cleavages of the receptor by ADAM metalloproteases and the γ-secretase complex. Ligand-induced proteolysis of NOTCH receptors liberates the NOTCH intracellular domain (NICD). The canonical signal links the NICD to the nuclear CSL (CBF-1, Suppressor of Hairless, Lag-1; RBPJ in vertebrates) transcriptional regulator [1], which is part of a repressor complex that includes NCoR/SMRT and histone deacetylases, and is bound to the promoters of target genes. Upon NICD binding, the composition of the RBPJ complex changes and co-activators including mastermind-like protein 1 (MAML-1) and histone acetyltransferases (p300, CBP) are recruited to initiate gene transcription [2]. NOTCH controls the expression of a wide range of target genes on the basis of the tissue and cell type that reflects its multiple functions. The basic helix-loop-helix (bHLH) transcription factors of the HES/HEY family were the first NOTCH target genes to be described [3]. They mainly function as transcriptional repressors in heteromeric complexes with Transducin-Like Enhancer-of-split (TLE) proteins, and their expression is in some cases finely tuned by autoregulatory loops that result in oscillatory expression crucial to regulate, for example, stem cell maintenance and cell fate decision during embryonic development of the nervous system [4]. An additional level of regulation, at least in the brain, is exerted by a class of dominant negative regulators of bHLH proteins, the inhibitor of DNA-binding (ID) proteins. IDs can form heterodimers with bHLH factors and, in most cases, counteract their activity [5,6,7,8]. NICD/RBPJ can regulate the activity of gene promoters and play a fundamental role at distant long-range enhancers [9,10], as has been described in T cell acute lymphoblastic leukemia (T-ALL) [11,12].
NOTCH signaling is highly conserved from invertebrates to humans and is involved in the regulation of a variety of cellular processes throughout life, including cell proliferation, stem cell maintenance, cell fate decisions, and differentiation. Accordingly, deregulation of NOTCH signaling occurs in several human diseases, including cancer [1]. Due to its pleiotropic functions, NOTCH signaling is implicated in many aspects of tumor development where it can act either as an oncogene or a tumor suppressor, as extensively reviewed by others [13,14]. The balance between one or the other role is determined by many factors, such as the cell type, the stage of tumor development, or the target genes involved [13,14]. Genetic alterations in the genes of different NOTCH pathway components that can lead to either enhanced or decreased NOTCH signaling have been identified in multiple cancers [13,14,15]. Mutations that directly occur in NOTCH receptor genes can alter different portions of the protein [15]. NOTCH activating mutations often occur in the negative regulatory region (NRR), which regulates receptor cleavage, or in the PEST domain, which is involved in NICD degradation, thus leading to ligand-independent or prolonged receptor activation [16,17,18,19,20]. In contrast, NOTCH inactivating mutations predominantly occur in the epidermal growth factor (EGF)-like repeats of the extracellular portion of the receptors. These mutations, including point mutations, focal deletions, and nonsense and missense mutations, are predicted to prevent ligand binding or generate a truncated protein and are often associated with downregulated expression of NOTCH target genes [21,22,23,24,25,26].
Historically, NOTCH signaling in cancer has been extensively studied in the context of T-ALL, where NOTCH acts as an oncogene and NOTCH-activating mutations are present in more than 50% of patients [17]. Different laboratories have identified translocations or mutation clusters within the NOTCH1 receptor gene that significantly favor tumor progression by causing a ligand-independent constitutive activation of the pathway [27]. Evidence that NOTCH can also promote tumor growth in solid tumors comes, for example, from breast cancer. It has been demonstrated that integration of the mouse mammary tumor virus (MMTV) causes rearrangement and activation of a particular locus containing the Notch4 sequence and this ultimately results in cancer development [28,29]. Recurrent gene rearrangements in NOTCH1 and NOTCH2, inducing NOTCH pathway activation, have also been identified in human estrogen receptor (ER)-negative and triple-negative breast cancer samples and cell lines [19]. In T-ALL and breast cancers, evidence points to c-Myc as an important effector of the tumor-promoting function of NOTCH [17,19,30]. However, NOTCH could also cooperate with known “cancer drivers” including breast cancer gene 1 (BRCA1), whose loss-of-function is responsible for the majority of breast cancers [31]. Miao and colleagues (2020) have recently suggested that NOTCH1 activation can regulate the cell cycle and attenuate BRCA1 deficiency-induced G2/M blockade and genomic instability, thus providing tumor cells with a survival advantage [31]. Moreover, it has been proposed that NOTCH signaling could promote epithelial-to-mesenchymal transition (EMT), enhancing tumor cell aggressiveness and metastatic potential [31,32].
The NOTCH pathway can also be tumor-suppressive in other contexts. The first evidence for a tumor suppressor role of NOTCH came from studies in mouse and human keratinocytes [33,34]. Later, NOTCH-inactivating mutations were identified in head and neck squamous cell carcinoma (HNSCC) [21,22] and small cell lung cancer (SCLC) [35,36,37]. In 2011, two independent whole-genome sequencing studies reported NOTCH-inactivating mutations in two different cohorts of patients with HNSCC [21,22]. These findings were further supported by a more recent and comprehensive genomic characterization of HNSCC that found NOTCH mutations in approximately 20% of the patients [38]. In SCLC, a highly aggressive and therapy-resistant lung cancer characterized by the expression of neuroendocrine (NE) markers, a recent in vivo clonal analysis demonstrated that a rare population of pulmonary NE stem cells could be induced to reactivate and differentiate upon injury and that this process is regulated by NOTCH. However, when NOTCH signaling is inhibited, the differentiation program is blocked and NE cells remain in a highly self-renewing state that is prone to transformation [39]. NOTCH-inactivating mutations correlating with a poorer patient prognosis have also been found in approximately 40% of patients with bladder cancer. Intriguingly, half of those patients did not carry other concomitant mutations in well-known oncogenic drivers including FGFR3 or RAS, suggesting a prominent role for NOTCH signaling in tumor initiation [23]. Moreover, complete or partial loss of chromosome 9, where the NOTCH1 gene is located (9q34.3), is a common chromosomal aberration in bladder carcinoma [40]. In these tumors, NOTCH acts as a suppressor of cell proliferation by upregulating multiple members of the dual-specific phosphatase (DUSP) family, which inhibit Extracellular signal-Regulated Kinase 1/2 (ERK1/2) phosphorylation. As a consequence, NOTCH mutant tumor cells show increased ERK1/2 phosphorylation that can be reverted by NOTCH activation [23].
Intriguingly, there is evidence that NOTCH signaling can play both tumor-promoting and tumor-suppressive roles, even within the same organ. For instance, although the growth of HNSCC is largely driven by NOTCH inactivation [41], occasional Notch gain-of-function mutations have been reported in oral squamous cell carcinoma (OSCC) [42]. In the hematopoietic system, while an oncogenic role of NOTCH has been described in both acute (T-ALL) and chronic (CLL) forms of lymphocytic leukemia [17,18], a tumor-suppressive role has been proposed in chronic myelomonocytic leukemia (CMML) [43] and also suggested in acute myeloid leukemia (AML) [44]. Such dualism has been linked to the function of NOTCH in the regulation of cell fate choices during immune cell development. Multiple in vitro and in vivo studies have demonstrated that NOTCH favors T cell over B cell commitment and myeloid differentiation [45,46,47]. Consequently, NOTCH gain-of-function mutations lead to a rapid and abnormal expansion of T cells at the expense of other cell lineages, whereas NOTCH inactivation, particularly in the stromal compartment, causes an increase in myeloid progenitors and granulocyte/macrophage descendants, resulting in myeloid hyperplasia and myeloproliferative-like disease [43,48,49,50,51]. Hence, depending on the cell type, NOTCH signaling can play opposite roles in the development of hematological malignancies. A dual role for NOTCH signaling is also evident in some solid tumors, including lung cancer. The most prevalent form of lung tumors is non-small-cell lung cancer (NSCLC), a heterogeneous group of neoplasms that includes lung adenocarcinoma and squamous cell lung carcinoma in which NOTCH signaling activity has been suggested to promote and suppress tumor growth, respectively [14,37,52,53,54]. However, perhaps the most emblematic example of the fascinating complexity of how the NOTCH pathway can orchestrate tumor development is given by SCLC, an infrequent but very aggressive subtype of lung cancer. Lim and colleagues (2017) proposed that NOTCH signaling can simultaneously be oncogenic and tumor-suppressive in different cell subpopulations of an individual tumor, although intratumoral heterogeneity generated by NOTCH activity promotes overall SCLC growth [55]. The authors described the presence of two “symbiotic” cell types: slowly-proliferating and chemoresistant non-NE cells and actively dividing NE cells. NOTCH activation triggers a non-NE fate switch that slows tumor growth but also gives rise to non-NE cells that sustain NE cell expansion by providing trophic support. Therefore, while NOTCH activation delays initial tumor progression, it can provide a survival advantage after chemotherapy and fuel rapid tumor relapse [55].

2. NOTCH Signaling in Neural Stem/Progenitor Cells and Glioma Formation

Gliomas account for approximately 30% of all brain tumors and around 80% of malignant brain tumors [56]. They can be generically divided into diffuse low- and intermediate-grade II–III gliomas (low-grade gliomas, LGGs) and the most aggressive grade IV gliomas or glioblastoma multiforme (GBM). Although patients with LGGs have a more favorable prognosis, the diffuse infiltrative nature of LGGs makes a complete neurosurgical resection almost impossible, often resulting in recurrence and progression towards higher-grade gliomas [24]. Indeed, around 20% of GBMs are secondary tumors arising from preexisting LGGs [26]. Recent advances in omics techniques have allowed for the integration of clinical and histopathological data with a deep characterization of the genetic, epigenetic, expression, and metabolic profiles of gliomas [57]. The scenario that has emerged is that of a heterogeneous group of glioma subtypes, each characterized by different survival rates and responses to treatments. The importance of these findings has encouraged the World Health Organization to include molecular parameters together with the classical histological markers in the diagnosis of gliomas [58]. Given the broad heterogeneity within gliomas and the highly context-dependent roles of NOTCH signaling in different cancers even within the same tumor type, it is perhaps not surprising that the functions of NOTCH in glioma still remain controversial.
NOTCH activity is fundamental to maintain neural stem cell (NSC) and progenitor identity and to regulate cell fate decisions in the developing and adult brain [4]. Since uncontrolled proliferation and impaired differentiation of NSCs and glial progenitors may lead to glioma formation [59,60], understanding how these processes are regulated would be important to target cancer-initiating cells. After formation of the central nervous system is completed, NSCs persist within two restricted regions of the mammalian brain, the subgranular zone of the hippocampal dentate gyrus (DG) and the subventricular zone (SVZ) of the lateral ventricles [61,62]. Most adult NSCs are quiescent and only a proportion of them divide at any given time to self-renew and generate intermediate progenitors (IPs). IPs, in contrast, actively proliferate and differentiate mainly into neuroblasts, although glial cells can also be generated in the adult NSC niches [63,64,65]. The entire process is finely tuned by a variety of different mechanisms that ensure neuron generation and, at the same time, maintenance of a stem cell pool. Within the neurogenic lineage, the canonical RBPJ-mediated NOTCH pathway and the NOTCH target gene Hes5 are restricted to NSCs [66,67,68,69]. While there is evidence for adult stem and progenitor heterogeneity, NOTCH-dependence seems to be a common feature among NSC populations [62]. Indeed, both quiescent and actively proliferating NSCs express Hes5, indicating NOTCH pathway activation [70,71,72], and simultaneous Notch1 and Notch2 or Rbpj gene deletion in mice causes NSC proliferation, differentiation, and exhaustion [67,68,73]. Interestingly, however, and in contrast to global NOTCH pathway inhibition, blocking the function of individual NOTCH receptors differentially affects active versus quiescent NSC populations, which seem to preferentially rely on NOTCH1 and NOTCH2/3, respectively [73,74,75,76,77]. Recent data from our lab suggest that NOTCH2 can promote NSC quiescence by inducing the expression of ID4 [77]. ID4 and NOTCH signaling synergize to inhibit excessive accumulation of the proneural factor ASCL1 [77,78], which stimulates NSC proliferation and differentiation [79,80]. While NOTCH target genes of the HES/HEY family repress ASCL1 gene transcription [78], ID4 also facilitates the degradation of ASCL1 protein [81].
NOTCH signaling also regulates both the astrocyte and oligodendrocyte lineages. During embryonic development, NOTCH favors the gliogenic switch of radial glial cells by driving the expression of glial cell markers GFAP [82], BLBP [83], and NFIA [84]. The induction of an astrogliogenic program also requires NOTCH signaling after ischemia in the adult brain [85]. NOTCH signaling remains active in differentiated astrocytes but is modulated in response to injury. In striatal astrocytes, NOTCH signaling is reduced after stroke, resulting in ectopic glial proliferation and production of new neurons in non-neurogenic regions [86]. Reports indicate that NOTCH signaling also modulates the fate of oligodendrocyte progenitor cells (OPCs), another important cell of origin of glioma [60]. Jagged1-induced canonical NOTCH signaling restricts oligodendrocyte maturation and maintains OPCs during brain development and remyelination [87,88,89]. However, F3/Contactin can act as a non-canonical NOTCH ligand to promote OPC differentiation in a developmental stage specific manner [90].
Altogether, it is clear that NOTCH signaling plays a central role during homeostasis and injury response in NSCs and glial progenitors, both of which are potential cells-of-origin of brain tumors [60,91,92,93,94]. In line with this, NOTCH ligands, receptors, and downstream targets are expressed in several types of brain tumor [95,96,97,98,99,100] and varying levels of NOTCH activity can contribute to intra-tumor heterogeneity by promoting stem cell character in subpopulations of glioma cells [101,102]. Clearly, NOTCH signaling potentially regulates multiple steps of gliomagenesis, including tumor initiation, progression, and recurrence. Yet, a consensus on the role of NOTCH in glioma development is still missing.

3. NOTCH as an Oncogene in Glioma

Evidence indicates that NOTCH signaling can promote glioma aggressiveness in some contexts (Figure 1). The presence of self-renewing glioma stem cells (GSCs) that have increased DNA repair capacity and expression of ATP-binding cassette (ABC) multidrug transporters, and that differentiate into less-tumorigenic cancer cells that form the tumor bulk, is one phenomenon that can confer therapy resistance in glioma [101,103,104,105,106]. Reminiscent of its role in healthy NSCs, NOTCH can facilitate stem cell character in brain tumors and has therefore been considered a promising target for the development of more effective glioma therapies. Data from both human tumors and murine glioma models suggest that NOTCH signaling is preferentially active in subpopulations of glioma cells [101,102,107]. In agreement with this, in vitro and xenotransplantation studies with glioma cell lines have indicated that CD133-positive GSCs are particularly sensitive to γ-secretase inhibitors (GSI) or NOTCH1/2 knockdown compared to CD133-negative glioma cells [108,109]. Blocking NOTCH signaling or RBPJ reduced clonogenic potential in tumor-sphere assays and engraftment capacity in glioma xenograft models [108,109,110]. Conversely NICD overexpression, although not sufficient alone to induce brain tumors in the mouse brain [111,112], could increase cell survival due to radio-resistance and side population phenotype in glioma cells [101,108,109]. Combinatorial treatment of GSIs and radiation was more effective at inhibiting self-renewal than radiation alone [109,113], a synergistic effect that was partially mediated by NOTCH enhancing Akt and Stat3 phosphorylation [108,109]. NOTCH signaling could also promote a malignant phenotype in human glioma cell lines and xenograft models by repressing the expression of the promyelocytic leukemia protein (PML) tumor suppressor [114] and inducing the oncogenic long non-coding RNA TUG1 [115]. Reports have shown that knocking down the NOTCH ligands Jagged1 or Dll1 by RNA interference reduced survival and growth of tumor cells in multiple glioma cell lines [99,116] and high Jagged1 expression correlates with poor prognosis of glioma patients [116]. Interestingly, the extracellular matrix glycoprotein Tenascin-C and Jagged1 can reinforce each other’s expression, which could establish a feedback loop promoting tumor growth [116,117,118]. However, it is important to note that high levels of Jagged1 can also inhibit canonical NOTCH signaling in glioma cells, potentially through the activity of the Jagged1 intracellular domain [119].
In the adult SVZ neurogenic niche, NOTCH signaling plays a central role in maintaining the quiescent NSC pool [68,73], which is resistant to antimitotic treatment and can regenerate more actively proliferating progenitor cells [120,121]. Similarly, a relatively quiescent subset of chemotherapy-resistant glioma cells can propagate tumor growth after temozolomide administration in mouse models of GBM [104]. In line with this finding, a recent report demonstrated that pharmacological inhibition of receptor tyrosine kinases (RTK) prompts the emergence of slow-cycling and drug-tolerant persister cells in a subset of PDGFRA-amplified human glioma cell lines, and that an increase in NOTCH signaling activity allows transition to the persister state [122]. Although resistance to RTK inhibition could also develop independent of NOTCH in some cell clones [123], these data indicate that NOTCH can contribute to cell plasticity and drug tolerance in glioma by virtue of its stem cell-promoting activity. In this context, NOTCH inhibition could be exploited to release high expression of the proneural transcription factor ASCL1 in a subset of GSCs and induce their terminal neuronal differentiation [124,125].
Adult SVZ NSCs reside in a specialized vascular niche [126,127] that can foster stem cell character and repress differentiation through endothelial-derived factors that positively modulate NOTCH-dependent transcription [66,128]. In analogy to the SVZ niche, it has been proposed that GSCs reside in the proximity of blood vessels and are exposed to factors produced by endothelial cells [129]. Among these, nitric oxide has been shown to activate NOTCH signaling and promote stem-like character in PDGF-driven glioma [101]. Interestingly, NOTCH signaling is also augmented in hypoxic tumor regions by Vasorin-mediated stabilization of NICD, and hypoxia-induced expression of Vasorin promotes glioma aggressiveness [130]. Finally, there is evidence for dormant GSC populations residing at the invasive front of the tumor [131], where Jagged1 expressed by nerve fibers could facilitate NOTCH1+CD133+ glioma cell invasion of white matter tracts through a SOX9-SOX2-NOTCH1 feedback loop [132].

4. NOTCH as a Tumor Suppressor in Glioma

Surprisingly, recent data suggest that NOTCH signal inhibition may be an important molecular event in the formation of some forms of glioma (Figure 1). Genome-wide analyses in patients with LGG identified mutations in NOTCH signaling components in a significant proportion of isocitrate dehydrogenase (IDH) mutant tumors [24,25,26,133]. These mutations are particularly frequent in the genes encoding the NOTCH1 and NOTCH2 receptors but less common in NOTCH3/4, and occur at similar positions to those demonstrated experimentally to inactivate NOTCH1 function in epithelial cancers [24,25,26,133]. In addition, less frequent mutations in the RBPJ gene were detected in LGG samples and were mutually exclusive with NOTCH1 mutations [134]. Accordingly, the expression of some NOTCH target genes is downregulated in NOTCH/RBPJ mutant gliomas, indicating pathway inactivation [24,26,134]. Expression of the proliferation marker MKI67 was found to be augmented in NOTCH1/RBPJ mutant brain tumors with low HES/HEY expression levels [134], and NOTCH mutations were a high-risk factor associated with shorter patient survival [135], suggesting that NOTCH signaling inhibition contributes to increased glioma aggressiveness. Consistent with this, frequent NOTCH pathway alterations were also detected when comparing progressed tumors with their corresponding lower-grade counterparts [26]. Although occasional NOTCH1/2 mutations could also be detected in IDH mutant astrocytomas with TP53 inactivation [25], genetic inactivation of NOTCH pathway components mainly occurs in oligodendrogliomas with co-deletion of chromosome arms 1p and 19q [24,25]. This is intriguing, as the genes for NOTCH2, MIB2, and the NOTCH targets HES2-5 and HEYL are all located on chromosome arm 1p [136,137], lending support to the hypothesis that NOTCH signaling inhibition contributes to the initiation of tumors with oligodendroglial characteristics. Interestingly, genetic inhibition of NOTCH can promote OPC proliferation and expansion in mouse models [138], and excessive activation of quiescent OPC populations leads to malignant transformation [139].
In contrast to LGG, NOTCH mutations are very rare in GBM. However, it is worth noting that levels of NOTCH signaling activation substantially vary among GBM samples and subtypes [98,107], and hemizygous or homozygous deletions of the 1p36 locus, which could affect the MIB2 and HES2-5 genes, occur in a proportion of GBMs [140,141]. Although a number of genes have been proposed as 1p36 candidate tumor suppressors, it is unclear if NOTCH inhibition can contribute to cancer development after 1p36 loss.
Indirect evidence that NOTCH signaling could have tumor suppressor activity in glioma comes from studies on the proneural transcription factor ASCL1, whose expression is normally repressed by canonical NOTCH targets of the HES/HEY family in NSCs and GSCs [78,107,124,142]. Lineage tracing experiments in mice have shown that ASCL1+ neural progenitors can be cells of origin of GBM [59,91] and upregulation of ASCL1 and inhibition of NOTCH signaling characterize astrocytoma progression [143]. ASCL1 expression is maintained in both xenografts from human proneural GBM samples and GBM mouse models, and ASCL1 can induce cell cycle genes and oncogenes, thereby promoting glioma cell proliferation in some contexts [122,144,145,146]. Accordingly, ASCL1 knockdown or conditional gene knockout prolong survival of glioma-bearing mice [145,146]. In contrast, high expression of a curated NOTCH-signaling gene set, including HES/HEY transcriptional repressors, correlates with less proliferative glioma cell subpopulations with putative GSC character [122]. These data are in line with the role of NOTCH signaling in reducing ASCL1 expression and preserving quiescence in NSCs [68,78,79,86,107].
More direct evidence that NOTCH signaling could restrict glioma formation in some contexts comes from studies taking advantage of genetic conditional gene deletion or overexpression approaches in vivo. NOTCH2 overexpression can inhibit glioma formation in mouse glioma models [107] and HEY2 overexpression can reduce the proliferation of murine and human glioma cells [147]. Conversely, ID2-mediated repression of HEY1 promotes NSC transformation, possibly by releasing the expression of OLIG2 [148], an oligodendroglial lineage determinant involved in glioma cell proliferation [149]. Simultaneous genetic deletion of Notch1 and Notch2 or Rbpj accelerates the growth of PDGF-driven GBMs in mice [107]. Moreover, inactivation of the Rbpj gene together with the tumor suppressor gene Trp53 induces aggressive, de novo forebrain tumors with primitive neuroectodermal features [107]. These findings are in line with an unbiased in vivo CRISPR screen that recently identified NOTCH1 among potential tumor suppressors in GBM [150] and with the observation that high expression of specific NOTCH target genes positively correlates with a better prognosis in defined subsets of LGGs and GBMs in humans [107].

5. Open Questions and Perspectives

NOTCH receptors and their downstream targets are potential candidates for specific drug targeting, and various strategies to modulate NOTCH for cancer therapy are being actively pursued [151]. However, understanding the molecular basis of NOTCH oncogenic and tumor-suppressive functions is fundamental in order to develop effective strategies to therapeutically modulate NOTCH in glioma. Although clinical trials with GSIs in patients with glioma have been reported, positive effects have only been seen in one study [151,152]. Given the broad spectrum of gamma-secretase target proteins and the dose-limiting toxicities of GSIs, it would be interesting to test the effects of small molecules that directly target the NOTCH transcriptional activation complex in glioma [153,154,155]. Considering the extensive intertumoral heterogeneity in glioma, it is possible that NOTCH inhibition would be beneficial only in a proportion of molecularly selected patients in a personalized therapy. For instance, studies in vitro suggest that PTEN and TP53 status may affect sensitivity to GSIs in GBM [156,157]. NOTCH and EGF receptor pathways can potentiate each other in glioma cells [158,159,160] and, therefore, the oncogenic function of NOTCH signaling could be more apparent in primary GBMs of the classical subtype [161] than in GBMs with OPC-like proneural features [162] or IDH mutant LGGs [24,25]. The precise role of NOTCH in different forms of glioma requires further studies.
To some extent, the strength of NOTCH signaling might also explain the reported discrepancies in the oncogenic versus tumor-suppressive functions of NOTCH in the brain. NOTCH signaling is required to maintain a low and oscillatory ASCL1 expression in order to promote NSC self-renewal [6,78]. However, sustained NOTCH signaling activation induces proliferative quiescence of NSCs [78]. Conversely, NOTCH inhibition induces activation of latent parenchymal progenitors [86] and NSC hyperproliferation before resulting in neuronal differentiation and NSC depletion [78]. Reminiscent of this, NOTCH inhibition could induce terminal differentiation of glioma cells in some contexts [124] but facilitate their proliferation in others [107]. NOTCH inhibition can contribute to the initiation of epithelial cancers by favoring the expansion of NOTCH mutant clones at the expense of wild-type cells [163]. Since analogous cell competition mechanisms play important roles during cancer progression [164], it would be interesting to determine if NOTCH-regulated competitive interactions occur between adjacent neural progenitors and if NOTCH mutations can contribute to glioma formation in this context. Reversible NOTCH signal activation likely plays a role in regulating glioma cell behavior during tumor development, but NOTCH-independent clones may arise under prolonged drug selection pressure [123]. Hence, it will also be important to address if the role of NOTCH in glioma varies on the basis of the stage of disease progression, as suggested for SCLC [55].
Finally, indication for a tumor-suppressive activity of the NOTCH pathway in glioma predominantly comes from an in vivo immunocompetent setting, and a recent CRISPR screen identified frequent co-mutation of the NOTCH1 receptor and B2m, an essential component of the MHC-I antigen presentation complex [150]. Interestingly, findings indicate that the immune response to cancer in the brain is shaped by the cancer type [165,166]. Whether NOTCH activity in tumor cells can regulate interactions with the glioma microenvironment and immune evasion remains unexplored.

6. Concluding Remarks

NOTCH signaling can act as an oncogene or a tumor suppressor, depending on the context [1]. Although it is clear that NOTCH plays central roles in glioma, its precise function has remained puzzling. The data reviewed here support the hypothesis of a dual role of the NOTCH pathway as an oncogene and a tumor suppressor in glioma (Figure 1), similar to what has been suggested in some other malignancies. Human gliomas comprise multiple disease subtypes that differ at the genetic, epigenetic, and transcriptional levels, and this intertumoral heterogeneity could be one critical factor underlying the observed discrepancies in NOTCH function. In addition, differences in the outcome of NOTCH modulation likely relate to the stage of disease progression, crosstalk with other signaling pathways, and intratumoral (stem) cell heterogeneity. While diverse molecular bases of NOTCH oncogenic function in glioma have been addressed in previous studies, mechanistic data on the NOTCH tumor-suppressive activity in brain tumor subtypes are still lacking. It is also unclear if NOTCH-regulated cell competition mechanisms are in place during glioma initiation and progression. Finally, whether NOTCH activity in glioma cells regulates the crosstalk with the tumor microenvironment and immune cells in particular remains unexplored. Clearly, the multiple tasks of NOTCH signaling in glioma deserve further scrutiny.

Author Contributions

All authors contributed to manuscript conceptualization and preparation. All authors have read and agreed to the published version of the manuscript.

Funding

Swiss Cancer League (KLS-4518-08-2018), Stiftung zur Krebsbekämpfung, Huggenberger-Bischoff Stiftung zur Krebsforschung, Stiftung für krebskranke Kinder Regio Basiliensis, and Forschungsfonds Nachwuchsforschende of the University of Basel.

Acknowledgments

We thank the members of the Taylor lab for critical reading of the manuscript and for helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siebel, C.; Lendahl, U. Notch Signaling in Development, Tissue Homeostasis, and Disease. Physiol. Rev. 2017, 97, 1235–1294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kopan, R.; Ilagan, M.X. The canonical Notch signaling pathway: Unfolding the activation mechanism. Cell 2009, 137, 216–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Fischer, A.; Gessler, M. Delta-Notch–and then? Protein interactions and proposed modes of repression by Hes and Hey bHLH factors. Nucleic. Acids. Res. 2007, 35, 4583–4596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Pierfelice, T.; Alberi, L.; Gaiano, N. Notch in the vertebrate nervous system: An old dog with new tricks. Neuron 2011, 69, 840–855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Boareto, M.; Iber, D.; Taylor, V. Differential interactions between Notch and ID factors control neurogenesis by modulating Hes factor autoregulation. Development 2017, 144, 3465–3474. [Google Scholar] [CrossRef] [Green Version]
  6. Imayoshi, I.; Isomura, A.; Harima, Y.; Kawaguchi, K.; Kori, H.; Miyachi, H.; Fujiwara, T.; Ishidate, F.; Kageyama, R. Oscillatory control of factors determining multipotency and fate in mouse neural progenitors. Science 2013, 342, 1203–1208. [Google Scholar] [CrossRef] [Green Version]
  7. Bai, G.; Sheng, N.; Xie, Z.; Bian, W.; Yokota, Y.; Benezra, R.; Kageyama, R.; Guillemot, F.; Jing, N. Id sustains Hes1 expression to inhibit precocious neurogenesis by releasing negative autoregulation of Hes1. Dev. Cell 2007, 13, 283–297. [Google Scholar] [CrossRef] [Green Version]
  8. Nam, H.S.; Benezra, R. High levels of Id1 expression define B1 type adult neural stem cells. Cell Stem. Cell 2009, 5, 515–526. [Google Scholar] [CrossRef] [Green Version]
  9. Castel, D.; Mourikis, P.; Bartels, S.J.; Brinkman, A.B.; Tajbakhsh, S.; Stunnenberg, H.G. Dynamic binding of RBPJ is determined by Notch signaling status. Genes Dev. 2013, 27, 1059–1071. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, H.; Zou, J.; Zhao, B.; Johannsen, E.; Ashworth, T.; Wong, H.; Pear, W.S.; Schug, J.; Blacklow, S.C.; Arnett, K.L.; et al. Genome-wide analysis reveals conserved and divergent features of Notch1/RBPJ binding in human and murine T-lymphoblastic leukemia cells. Proc. Natl. Acad. Sci. USA 2011, 108, 14908–14913. [Google Scholar] [CrossRef] [Green Version]
  11. Herranz, D.; Ambesi-Impiombato, A.; Palomero, T.; Schnell, S.A.; Belver, L.; Wendorff, A.A.; Xu, L.; Castillo-Martin, M.; Llobet-Navas, D.; Cordon-Cardo, C.; et al. A NOTCH1-driven MYC enhancer promotes T cell development, transformation and acute lymphoblastic leukemia. Nat. Med. 2014, 20, 1130–1137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Yashiro-Ohtani, Y.; Wang, H.; Zang, C.; Arnett, K.L.; Bailis, W.; Ho, Y.; Knoechel, B.; Lanauze, C.; Louis, L.; Forsyth, K.S.; et al. Long-range enhancer activity determines Myc sensitivity to Notch inhibitors in T cell leukemia. Proc. Natl. Acad. Sci. USA 2014, 111, E4946–E4953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Nowell, C.S.; Radtke, F. Notch as a tumour suppressor. Nat. Rev. Cancer 2017, 17, 145–159. [Google Scholar] [CrossRef] [PubMed]
  14. Ntziachristos, P.; Lim, J.S.; Sage, J.; Aifantis, I. From fly wings to targeted cancer therapies: A centennial for notch signaling. Cancer Cell 2014, 25, 318–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Egloff, A.M.; Grandis, J.R. Molecular pathways: Context-dependent approaches to Notch targeting as cancer therapy. Clin. Cancer Res. 2012, 18, 5188–5195. [Google Scholar] [CrossRef] [Green Version]
  16. Ellisen, L.W.; Bird, J.; West, D.C.; Soreng, A.L.; Reynolds, T.C.; Smith, S.D.; Sklar, J. TAN-1, the human homolog of the Drosophila notch gene, is broken by chromosomal translocations in T lymphoblastic neoplasms. Cell 1991, 66, 649–661. [Google Scholar] [CrossRef]
  17. Weng, A.P.; Ferrando, A.A.; Lee, W.; Morris, J.P.T.; Silverman, L.B.; Sanchez-Irizarry, C.; Blacklow, S.C.; Look, A.T.; Aster, J.C. Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia. Science 2004, 306, 269–271. [Google Scholar] [CrossRef] [Green Version]
  18. Puente, X.S.; Pinyol, M.; Quesada, V.; Conde, L.; Ordonez, G.R.; Villamor, N.; Escaramis, G.; Jares, P.; Bea, S.; Gonzalez-Diaz, M.; et al. Whole-genome sequencing identifies recurrent mutations in chronic lymphocytic leukaemia. Nature 2011, 475, 101–105. [Google Scholar] [CrossRef] [Green Version]
  19. Robinson, D.R.; Kalyana-Sundaram, S.; Wu, Y.M.; Shankar, S.; Cao, X.; Ateeq, B.; Asangani, I.A.; Iyer, M.; Maher, C.A.; Grasso, C.S.; et al. Functionally recurrent rearrangements of the MAST kinase and Notch gene families in breast cancer. Nat. Med. 2011, 17, 1646–1651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Wang, K.; Zhang, Q.; Li, D.; Ching, K.; Zhang, C.; Zheng, X.; Ozeck, M.; Shi, S.; Li, X.; Wang, H.; et al. PEST domain mutations in Notch receptors comprise an oncogenic driver segment in triple-negative breast cancer sensitive to a gamma-secretase inhibitor. Clin. Cancer Res. 2015, 21, 1487–1496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Agrawal, N.; Frederick, M.J.; Pickering, C.R.; Bettegowda, C.; Chang, K.; Li, R.J.; Fakhry, C.; Xie, T.X.; Zhang, J.; Wang, J.; et al. Exome sequencing of head and neck squamous cell carcinoma reveals inactivating mutations in NOTCH1. Science 2011, 333, 1154–1157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Stransky, N.; Egloff, A.M.; Tward, A.D.; Kostic, A.D.; Cibulskis, K.; Sivachenko, A.; Kryukov, G.V.; Lawrence, M.S.; Sougnez, C.; McKenna, A.; et al. The mutational landscape of head and neck squamous cell carcinoma. Science 2011, 333, 1157–1160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Rampias, T.; Vgenopoulou, P.; Avgeris, M.; Polyzos, A.; Stravodimos, K.; Valavanis, C.; Scorilas, A.; Klinakis, A. A new tumor suppressor role for the Notch pathway in bladder cancer. Nat. Med. 2014, 20, 1199–1205. [Google Scholar] [CrossRef] [PubMed]
  24. Brat, D.J.; Verhaak, R.G.; Aldape, K.D.; Yung, W.K.; Salama, S.R.; Cooper, L.A.; Rheinbay, E.; Miller, C.R.; Vitucci, M.; Morozova, O.; et al. Comprehensive, Integrative Genomic Analysis of Diffuse Lower-Grade Gliomas. N. Engl. J. Med. 2015, 372, 2481–2498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Suzuki, H.; Aoki, K.; Chiba, K.; Sato, Y.; Shiozawa, Y.; Shiraishi, Y.; Shimamura, T.; Niida, A.; Motomura, K.; Ohka, F.; et al. Mutational landscape and clonal architecture in grade II and III gliomas. Nat. Genet. 2015, 47, 458–468. [Google Scholar] [CrossRef] [PubMed]
  26. Bai, H.; Harmanci, A.S.; Erson-Omay, E.Z.; Li, J.; Coskun, S.; Simon, M.; Krischek, B.; Ozduman, K.; Omay, S.B.; Sorensen, E.A.; et al. Integrated genomic characterization of IDH1-mutant glioma malignant progression. Nat. Genet. 2016, 48, 59–66. [Google Scholar] [CrossRef] [Green Version]
  27. Radtke, F.; Fasnacht, N.; Macdonald, H.R. Notch signaling in the immune system. Immunity 2010, 32, 14–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Jhappan, C.; Gallahan, D.; Stahle, C.; Chu, E.; Smith, G.H.; Merlino, G.; Callahan, R. Expression of an activated Notch-related int-3 transgene interferes with cell differentiation and induces neoplastic transformation in mammary and salivary glands. Genes Dev. 1992, 6, 345–355. [Google Scholar] [CrossRef] [Green Version]
  29. Robbins, J.; Blondel, B.J.; Gallahan, D.; Callahan, R. Mouse mammary tumor gene int-3: A member of the notch gene family transforms mammary epithelial cells. J. Virol. 1992, 66, 2594–2599. [Google Scholar] [CrossRef] [Green Version]
  30. Klinakis, A.; Szabolcs, M.; Politi, K.; Kiaris, H.; Artavanis-Tsakonas, S.; Efstratiadis, A. Myc is a Notch1 transcriptional target and a requisite for Notch1-induced mammary tumorigenesis in mice. Proc. Natl. Acad. Sci. USA 2006, 103, 9262–9267. [Google Scholar] [CrossRef] [Green Version]
  31. Miao, K.; Lei, J.H.; Valecha, M.V.; Zhang, A.; Xu, J.; Wang, L.; Lyu, X.; Chen, S.; Miao, Z.; Zhang, X.; et al. NOTCH1 activation compensates BRCA1 deficiency and promotes triple-negative breast cancer formation. Nat. Commun. 2020, 11, 3256. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Z.; Li, Y.; Kong, D.; Sarkar, F.H. The role of Notch signaling pathway in epithelial-mesenchymal transition (EMT) during development and tumor aggressiveness. Curr. Drug Targets 2010, 11, 745–751. [Google Scholar] [CrossRef]
  33. Lefort, K.; Mandinova, A.; Ostano, P.; Kolev, V.; Calpini, V.; Kolfschoten, I.; Devgan, V.; Lieb, J.; Raffoul, W.; Hohl, D.; et al. Notch1 is a p53 target gene involved in human keratinocyte tumor suppression through negative regulation of ROCK1/2 and MRCKalpha kinases. Genes Dev. 2007, 21, 562–577. [Google Scholar] [CrossRef] [Green Version]
  34. Nicolas, M.; Wolfer, A.; Raj, K.; Kummer, J.A.; Mill, P.; van Noort, M.; Hui, C.C.; Clevers, H.; Dotto, G.P.; Radtke, F. Notch1 functions as a tumor suppressor in mouse skin. Nat. Genet. 2003, 33, 416–421. [Google Scholar] [CrossRef]
  35. George, J.; Lim, J.S.; Jang, S.J.; Cun, Y.; Ozretic, L.; Kong, G.; Leenders, F.; Lu, X.; Fernandez-Cuesta, L.; Bosco, G.; et al. Comprehensive genomic profiles of small cell lung cancer. Nature 2015, 524, 47–53. [Google Scholar] [CrossRef] [PubMed]
  36. Sriuranpong, V.; Borges, M.W.; Ravi, R.K.; Arnold, D.R.; Nelkin, B.D.; Baylin, S.B.; Ball, D.W. Notch signaling induces cell cycle arrest in small cell lung cancer cells. Cancer Res. 2001, 61, 3200–3205. [Google Scholar] [PubMed]
  37. 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] [Green Version]
  38. Cancer Genome Atlas Network. Comprehensive genomic characterization of head and neck squamous cell carcinomas. Nature 2015, 517, 576–582. [Google Scholar] [CrossRef] [Green Version]
  39. Ouadah, Y.; Rojas, E.R.; Riordan, D.P.; Capostagno, S.; Kuo, C.S.; Krasnow, M.A. Rare Pulmonary Neuroendocrine Cells Are Stem Cells Regulated by Rb, p53, and Notch. Cell 2019, 179, 403–416.e23. [Google Scholar] [CrossRef] [Green Version]
  40. Kimura, F.; Florl, A.R.; Seifert, H.H.; Louhelainen, J.; Maas, S.; Knowles, M.A.; Schulz, W.A. Destabilization of chromosome 9 in transitional cell carcinoma of the urinary bladder. Br. J. Cancer 2001, 85, 1887–1893. [Google Scholar] [CrossRef]
  41. 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]
  42. Song, X.; Xia, R.; Li, J.; Long, Z.; Ren, H.; Chen, W.; Mao, L. Common and complex Notch1 mutations in Chinese oral squamous cell carcinoma. Clin. Cancer Res. 2014, 20, 701–710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Klinakis, A.; Lobry, C.; Abdel-Wahab, O.; Oh, P.; Haeno, H.; Buonamici, S.; van De Walle, I.; Cathelin, S.; Trimarchi, T.; Araldi, E.; et al. A novel tumour-suppressor function for the Notch pathway in myeloid leukaemia. Nature 2011, 473, 230–233. [Google Scholar] [CrossRef] [PubMed]
  44. Lobry, C.; Ntziachristos, P.; Ndiaye-Lobry, D.; Oh, P.; Cimmino, L.; Zhu, N.; Araldi, E.; Hu, W.; Freund, J.; Abdel-Wahab, O.; et al. Notch pathway activation targets AML-initiating cell homeostasis and differentiation. J. Exp. Med. 2013, 210, 301–319. [Google Scholar] [CrossRef] [PubMed]
  45. Bell, J.J.; Bhandoola, A. The earliest thymic progenitors for T cells possess myeloid lineage potential. Nature 2008, 452, 764–767. [Google Scholar] [CrossRef]
  46. Feyerabend, T.B.; Terszowski, G.; Tietz, A.; Blum, C.; Luche, H.; Gossler, A.; Gale, N.W.; Radtke, F.; Fehling, H.J.; Rodewald, H.R. Deletion of Notch1 converts pro-T cells to dendritic cells and promotes thymic B cells by cell-extrinsic and cell-intrinsic mechanisms. Immunity 2009, 30, 67–79. [Google Scholar] [CrossRef] [Green Version]
  47. Wada, H.; Masuda, K.; Satoh, R.; Kakugawa, K.; Ikawa, T.; Katsura, Y.; Kawamoto, H. Adult T-cell progenitors retain myeloid potential. Nature 2008, 452, 768–772. [Google Scholar] [CrossRef]
  48. Kim, Y.W.; Koo, B.K.; Jeong, H.W.; Yoon, M.J.; Song, R.; Shin, J.; Jeong, D.C.; Kim, S.H.; Kong, Y.Y. Defective Notch activation in microenvironment leads to myeloproliferative disease. Blood 2008, 112, 4628–4638. [Google Scholar] [CrossRef]
  49. Wang, L.; Zhang, H.; Rodriguez, S.; Cao, L.; Parish, J.; Mumaw, C.; Zollman, A.; Kamoka, M.M.; Mu, J.; Chen, D.Z.; et al. Notch-dependent repression of miR-155 in the bone marrow niche regulates hematopoiesis in an NF-kappaB-dependent manner. Cell Stem. Cell 2014, 15, 51–65. [Google Scholar] [CrossRef] [Green Version]
  50. Yao, D.; Huang, Y.; Huang, X.; Wang, W.; Yan, Q.; Wei, L.; Xin, W.; Gerson, S.; Stanley, P.; Lowe, J.B.; et al. Protein O-fucosyltransferase 1 (Pofut1) regulates lymphoid and myeloid homeostasis through modulation of Notch receptor ligand interactions. Blood 2011, 117, 5652–5662. [Google Scholar] [CrossRef] [Green Version]
  51. Yoda, M.; Kimura, T.; Tohmonda, T.; Uchikawa, S.; Koba, T.; Takito, J.; Morioka, H.; Matsumoto, M.; Link, D.C.; Chiba, K.; et al. Dual functions of cell-autonomous and non-cell-autonomous ADAM10 activity in granulopoiesis. Blood 2011, 118, 6939–6942. [Google Scholar] [CrossRef] [PubMed]
  52. Licciulli, S.; Avila, J.L.; Hanlon, L.; Troutman, S.; Cesaroni, M.; Kota, S.; Keith, B.; Simon, M.C.; Puré, E.; Radtke, F.; et al. Notch1 is required for Kras-induced lung adenocarcinoma and controls tumor cell survival via p53. Cancer Res. 2013, 73, 5974–5984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Westhoff, B.; Colaluca, I.N.; D’Ario, G.; Donzelli, M.; Tosoni, D.; Volorio, S.; Pelosi, G.; Spaggiari, L.; Mazzarol, G.; Viale, G.; et al. Alterations of the Notch pathway in lung cancer. Proc. Natl. Acad. Sci. USA 2009, 106, 22293–22298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Zheng, Y.; de la Cruz, C.C.; Sayles, L.C.; Alleyne-Chin, C.; Vaka, D.; Knaak, T.D.; Bigos, M.; Xu, Y.; Hoang, C.D.; Shrager, J.B.; et al. A rare population of CD24(+)ITGB4(+)Notch(hi) cells drives tumor propagation in NSCLC and requires Notch3 for self-renewal. Cancer Cell 2013, 24, 59–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Lim, J.S.; Ibaseta, A.; Fischer, M.M.; Cancilla, B.; O’Young, G.; Cristea, S.; Luca, V.C.; Yang, D.; Jahchan, N.S.; Hamard, C.; et al. Intratumoural heterogeneity generated by Notch signalling promotes small-cell lung cancer. Nature 2017, 545, 360–364. [Google Scholar] [CrossRef] [PubMed]
  56. Ostrom, Q.T.; Bauchet, L.; Davis, F.G.; Deltour, I.; Fisher, J.L.; Langer, C.E.; Pekmezci, M.; Schwartzbaum, J.A.; Turner, M.C.; Walsh, K.M.; et al. The epidemiology of glioma in adults: A “state of the science” review. Neuro. Oncol. 2014, 16, 896–913. [Google Scholar] [CrossRef] [Green Version]
  57. Reifenberger, G.; Wirsching, H.G.; Knobbe-Thomsen, C.B.; Weller, M. Advances in the molecular genetics of gliomas—Implications for classification and therapy. Nat. Rev. Clin. Oncol. 2017, 14, 434–452. [Google Scholar] [CrossRef] [PubMed]
  58. Louis, D.N.; Perry, A.; Reifenberger, G.; von Deimling, A.; Figarella-Branger, D.; Cavenee, W.K.; Ohgaki, H.; Wiestler, O.D.; Kleihues, P.; Ellison, D.W. The 2016 World Health Organization Classification of Tumors of the Central Nervous System: A summary. Acta Neuropathol. 2016, 131, 803–820. [Google Scholar] [CrossRef] [Green Version]
  59. Alcantara Llaguno, S.; Sun, D.; Pedraza, A.M.; Vera, E.; Wang, Z.; Burns, D.K.; Parada, L.F. Cell-of-origin susceptibility to glioblastoma formation declines with neural lineage restriction. Nat. Neurosci. 2019, 22, 545–555. [Google Scholar] [CrossRef]
  60. Liu, C.; Sage, J.C.; Miller, M.R.; Verhaak, R.G.; Hippenmeyer, S.; Vogel, H.; Foreman, O.; Bronson, R.T.; Nishiyama, A.; Luo, L.; et al. Mosaic analysis with double markers reveals tumor cell of origin in glioma. Cell 2011, 146, 209–221. [Google Scholar] [CrossRef] [Green Version]
  61. Bond, A.M.; Ming, G.L.; Song, H. Adult Mammalian Neural Stem Cells and Neurogenesis: Five Decades Later. Cell Stem. Cell 2015, 17, 385–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Giachino, C.; Taylor, V. Notching up neural stem cell homogeneity in homeostasis and disease. Front. Neurosci. 2014, 8, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Gage, F.H.; Temple, S. Neural stem cells: Generating and regenerating the brain. Neuron 2013, 80, 588–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Gotz, M.; Nakafuku, M.; Petrik, D. Neurogenesis in the Developing and Adult Brain-Similarities and Key Differences. Cold Spring Harb. Perspect Biol. 2016, 8. [Google Scholar] [CrossRef] [Green Version]
  65. Navarro Negredo, P.; Yeo, R.W.; Brunet, A. Aging and Rejuvenation of Neural Stem Cells and Their Niches. Cell Stem. Cell 2020, 27, 202–223. [Google Scholar] [CrossRef]
  66. Andreu-Agullo, C.; Morante-Redolat, J.M.; Delgado, A.C.; Farinas, I. Vascular niche factor PEDF modulates Notch-dependent stemness in the adult subependymal zone. Nat. Neurosci. 2009, 12, 1514–1523. [Google Scholar] [CrossRef]
  67. Ehm, O.; Goritz, C.; Covic, M.; Schaffner, I.; Schwarz, T.J.; Karaca, E.; Kempkes, B.; Kremmer, E.; Pfrieger, F.W.; Espinosa, L.; et al. RBPJkappa-dependent signaling is essential for long-term maintenance of neural stem cells in the adult hippocampus. J. Neurosci. 2010, 30, 13794–13807. [Google Scholar] [CrossRef] [Green Version]
  68. Imayoshi, I.; Sakamoto, M.; Yamaguchi, M.; Mori, K.; Kageyama, R. Essential roles of Notch signaling in maintenance of neural stem cells in developing and adult brains. J. Neurosci. 2010, 30, 3489–3498. [Google Scholar] [CrossRef] [Green Version]
  69. Mizutani, K.; Yoon, K.; Dang, L.; Tokunaga, A.; Gaiano, N. Differential Notch signalling distinguishes neural stem cells from intermediate progenitors. Nature 2007, 449, 351–355. [Google Scholar] [CrossRef]
  70. Lugert, S.; Basak, O.; Knuckles, P.; Haussler, U.; Fabel, K.; Gotz, M.; Haas, C.A.; Kempermann, G.; Taylor, V.; Giachino, C. Quiescent and active hippocampal neural stem cells with distinct morphologies respond selectively to physiological and pathological stimuli and aging. Cell Stem. Cell 2010, 6, 445–456. [Google Scholar] [CrossRef] [Green Version]
  71. Lugert, S.; Vogt, M.; Tchorz, J.S.; Muller, M.; Giachino, C.; Taylor, V. Homeostatic neurogenesis in the adult hippocampus does not involve amplification of Ascl1(high) intermediate progenitors. Nat. Commun. 2012, 3, 670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Giachino, C.; Basak, O.; Lugert, S.; Knuckles, P.; Obernier, K.; Fiorelli, R.; Frank, S.; Raineteau, O.; Alvarez-Buylla, A.; Taylor, V. Molecular diversity subdivides the adult forebrain neural stem cell population. Stem. Cells 2014, 32, 70–84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Engler, A.; Rolando, C.; Giachino, C.; Saotome, I.; Erni, A.; Brien, C.; Zhang, R.; Zimber-Strobl, U.; Radtke, F.; Artavanis-Tsakonas, S.; et al. Notch2 Signaling Maintains NSC Quiescence in the Murine Ventricular-Subventricular Zone. Cell Rep. 2018, 22, 992–1002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Ables, J.L.; Decarolis, N.A.; Johnson, M.A.; Rivera, P.D.; Gao, Z.; Cooper, D.C.; Radtke, F.; Hsieh, J.; Eisch, A.J. Notch1 is required for maintenance of the reservoir of adult hippocampal stem cells. J. Neurosci. 2010, 30, 10484–10492. [Google Scholar] [CrossRef] [Green Version]
  75. Basak, O.; Giachino, C.; Fiorini, E.; Macdonald, H.R.; Taylor, V. Neurogenic subventricular zone stem/progenitor cells are Notch1-dependent in their active but not quiescent state. J. Neurosci. 2012, 32, 5654–5666. [Google Scholar] [CrossRef] [Green Version]
  76. Kawai, H.; Kawaguchi, D.; Kuebrich, B.D.; Kitamoto, T.; Yamaguchi, M.; Gotoh, Y.; Furutachi, S. Area-Specific Regulation of Quiescent Neural Stem Cells by Notch3 in the Adult Mouse Subependymal Zone. J. Neurosci. 2017, 37, 11867–11880. [Google Scholar] [CrossRef]
  77. Zhang, R.; Boareto, M.; Engler, A.; Louvi, A.; Giachino, C.; Iber, D.; Taylor, V. Id4 Downstream of Notch2 Maintains Neural Stem Cell Quiescence in the Adult Hippocampus. Cell Rep. 2019, 28, 1485–1498.e6. [Google Scholar] [CrossRef]
  78. Sueda, R.; Imayoshi, I.; Harima, Y.; Kageyama, R. High Hes1 expression and resultant Ascl1 suppression regulate quiescent vs. active neural stem cells in the adult mouse brain. Genes Dev. 2019, 33, 511–523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Andersen, J.; Urban, N.; Achimastou, A.; Ito, A.; Simic, M.; Ullom, K.; Martynoga, B.; Lebel, M.; Goritz, C.; Frisen, J.; et al. A transcriptional mechanism integrating inputs from extracellular signals to activate hippocampal stem cells. Neuron 2014, 83, 1085–1097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Urban, N.; van den Berg, D.L.; Forget, A.; Andersen, J.; Demmers, J.A.; Hunt, C.; Ayrault, O.; Guillemot, F. Return to quiescence of mouse neural stem cells by degradation of a proactivation protein. Science 2016, 353, 292–295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Blomfield, I.M.; Rocamonde, B.; Masdeu, M.D.M.; Mulugeta, E.; Vaga, S.; van den Berg, D.L.; Huillard, E.; Guillemot, F.; Urban, N. Id4 promotes the elimination of the pro-activation factor Ascl1 to maintain quiescence of adult hippocampal stem cells. eLife 2019, 8. [Google Scholar] [CrossRef]
  82. Ge, W.; Martinowich, K.; Wu, X.; He, F.; Miyamoto, A.; Fan, G.; Weinmaster, G.; Sun, Y.E. Notch signaling promotes astrogliogenesis via direct CSL-mediated glial gene activation. J. Neurosci. Res. 2002, 69, 848–860. [Google Scholar] [CrossRef] [PubMed]
  83. Anthony, T.E.; Mason, H.A.; Gridley, T.; Fishell, G.; Heintz, N. Brain lipid-binding protein is a direct target of Notch signaling in radial glial cells. Genes Dev. 2005, 19, 1028–1033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Deneen, B.; Ho, R.; Lukaszewicz, A.; Hochstim, C.J.; Gronostajski, R.M.; Anderson, D.J. The transcription factor NFIA controls the onset of gliogenesis in the developing spinal cord. Neuron 2006, 52, 953–968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Benner, E.J.; Luciano, D.; Jo, R.; Abdi, K.; Paez-Gonzalez, P.; Sheng, H.; Warner, D.S.; Liu, C.; Eroglu, C.; Kuo, C.T. Protective astrogenesis from the SVZ niche after injury is controlled by Notch modulator Thbs4. Nature 2013, 497, 369–373. [Google Scholar] [CrossRef] [Green Version]
  86. Magnusson, J.P.; Goritz, C.; Tatarishvili, J.; Dias, D.O.; Smith, E.M.; Lindvall, O.; Kokaia, Z.; Frisen, J. A latent neurogenic program in astrocytes regulated by Notch signaling in the mouse. Science 2014, 346, 237–241. [Google Scholar] [CrossRef]
  87. Hammond, T.R.; Gadea, A.; Dupree, J.; Kerninon, C.; Nait-Oumesmar, B.; Aguirre, A.; Gallo, V. Astrocyte-derived endothelin-1 inhibits remyelination through notch activation. Neuron 2014, 81, 588–602. [Google Scholar] [CrossRef] [Green Version]
  88. John, G.R.; Shankar, S.L.; Shafit-Zagardo, B.; Massimi, A.; Lee, S.C.; Raine, C.S.; Brosnan, C.F. Multiple sclerosis: Re-expression of a developmental pathway that restricts oligodendrocyte maturation. Nat. Med. 2002, 8, 1115–1121. [Google Scholar] [CrossRef]
  89. Wang, S.; Sdrulla, A.D.; diSibio, G.; Bush, G.; Nofziger, D.; Hicks, C.; Weinmaster, G.; Barres, B.A. Notch receptor activation inhibits oligodendrocyte differentiation. Neuron 1998, 21, 63–75. [Google Scholar] [CrossRef] [Green Version]
  90. Hu, Q.D.; Ang, B.T.; Karsak, M.; Hu, W.P.; Cui, X.Y.; Duka, T.; Takeda, Y.; Chia, W.; Sankar, N.; Ng, Y.K.; et al. F3/contactin acts as a functional ligand for Notch during oligodendrocyte maturation. Cell 2003, 115, 163–175. [Google Scholar] [CrossRef] [Green Version]
  91. Alcantara Llaguno, S.R.; Wang, Z.; Sun, D.; Chen, J.; Xu, J.; Kim, E.; Hatanpaa, K.J.; Raisanen, J.M.; Burns, D.K.; Johnson, J.E.; et al. Adult Lineage-Restricted CNS Progenitors Specify Distinct Glioblastoma Subtypes. Cancer Cell 2015, 28, 429–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Bachoo, R.M.; Maher, E.A.; Ligon, K.L.; Sharpless, N.E.; Chan, S.S.; You, M.J.; Tang, Y.; DeFrances, J.; Stover, E.; Weissleder, R.; et al. Epidermal growth factor receptor and Ink4a/Arf: Convergent mechanisms governing terminal differentiation and transformation along the neural stem cell to astrocyte axis. Cancer Cell 2002, 1, 269–277. [Google Scholar] [CrossRef] [Green Version]
  93. Holland, E.C.; Celestino, J.; Dai, C.; Schaefer, L.; Sawaya, R.E.; Fuller, G.N. Combined activation of Ras and Akt in neural progenitors induces glioblastoma formation in mice. Nat. Genet. 2000, 25, 55–57. [Google Scholar] [CrossRef] [PubMed]
  94. Lee, J.H.; Lee, J.E.; Kahng, J.Y.; Kim, S.H.; Park, J.S.; Yoon, S.J.; Um, J.Y.; Kim, W.K.; Lee, J.K.; Park, J.; et al. Human glioblastoma arises from subventricular zone cells with low-level driver mutations. Nature 2018, 560, 243–247. [Google Scholar] [CrossRef]
  95. Brennan, C.; Momota, H.; Hambardzumyan, D.; Ozawa, T.; Tandon, A.; Pedraza, A.; Holland, E. Glioblastoma subclasses can be defined by activity among signal transduction pathways and associated genomic alterations. PLoS ONE 2009, 4, e7752. [Google Scholar] [CrossRef]
  96. Cuevas, I.C.; Slocum, A.L.; Jun, P.; Costello, J.F.; Bollen, A.W.; Riggins, G.J.; McDermott, M.W.; Lal, A. Meningioma transcript profiles reveal deregulated Notch signaling pathway. Cancer Res. 2005, 65, 5070–5075. [Google Scholar] [CrossRef] [Green Version]
  97. Fan, X.; Mikolaenko, I.; Elhassan, I.; Ni, X.; Wang, Y.; Ball, D.; Brat, D.J.; Perry, A.; Eberhart, C.G. Notch1 and notch2 have opposite effects on embryonal brain tumor growth. Cancer Res. 2004, 64, 7787–7793. [Google Scholar] [CrossRef] [Green Version]
  98. Phillips, H.S.; Kharbanda, S.; Chen, R.; Forrest, W.F.; Soriano, R.H.; Wu, T.D.; Misra, A.; Nigro, J.M.; Colman, H.; Soroceanu, L.; et al. Molecular subclasses of high-grade glioma predict prognosis, delineate a pattern of disease progression, and resemble stages in neurogenesis. Cancer Cell 2006, 9, 157–173. [Google Scholar] [CrossRef] [Green Version]
  99. Purow, B.W.; Haque, R.M.; Noel, M.W.; Su, Q.; Burdick, M.J.; Lee, J.; Sundaresan, T.; Pastorino, S.; Park, J.K.; Mikolaenko, I.; et al. Expression of Notch-1 and its ligands, Delta-like-1 and Jagged-1, is critical for glioma cell survival and proliferation. Cancer Res. 2005, 65, 2353–2363. [Google Scholar] [CrossRef] [Green Version]
  100. Shih, A.H.; Holland, E.C. Notch signaling enhances nestin expression in gliomas. Neoplasia 2006, 8, 1072–1082. [Google Scholar] [CrossRef] [Green Version]
  101. Charles, N.; Ozawa, T.; Squatrito, M.; Bleau, A.M.; Brennan, C.W.; Hambardzumyan, D.; Holland, E.C. Perivascular nitric oxide activates notch signaling and promotes stem-like character in PDGF-induced glioma cells. Cell Stem. Cell 2010, 6, 141–152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Zhu, T.S.; Costello, M.A.; Talsma, C.E.; Flack, C.G.; Crowley, J.G.; Hamm, L.L.; He, X.; Hervey-Jumper, S.L.; Heth, J.A.; Muraszko, K.M.; et al. Endothelial cells create a stem cell niche in glioblastoma by providing NOTCH ligands that nurture self-renewal of cancer stem-like cells. Cancer Res. 2011, 71, 6061–6072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Bao, S.; Wu, Q.; McLendon, R.E.; Hao, Y.; Shi, Q.; Hjelmeland, A.B.; Dewhirst, M.W.; Bigner, D.D.; Rich, J.N. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature 2006, 444, 756–760. [Google Scholar] [CrossRef] [PubMed]
  104. Chen, J.; Li, Y.; Yu, T.S.; McKay, R.M.; Burns, D.K.; Kernie, S.G.; Parada, L.F. A restricted cell population propagates glioblastoma growth after chemotherapy. Nature 2012, 488, 522–526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Lathia, J.D.; Mack, S.C.; Mulkearns-Hubert, E.E.; Valentim, C.L.; Rich, J.N. Cancer stem cells in glioblastoma. Genes Dev. 2015, 29, 1203–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Singh, S.K.; Hawkins, C.; Clarke, I.D.; Squire, J.A.; Bayani, J.; Hide, T.; Henkelman, R.M.; Cusimano, M.D.; Dirks, P.B. Identification of human brain tumour initiating cells. Nature 2004, 432, 396–401. [Google Scholar] [CrossRef] [PubMed]
  107. Giachino, C.; Boulay, J.L.; Ivanek, R.; Alvarado, A.; Tostado, C.; Lugert, S.; Tchorz, J.; Coban, M.; Mariani, L.; Bettler, B.; et al. A Tumor Suppressor Function for Notch Signaling in Forebrain Tumor Subtypes. Cancer Cell 2015, 28, 730–742. [Google Scholar] [CrossRef] [Green Version]
  108. Fan, X.; Khaki, L.; Zhu, T.S.; Soules, M.E.; Talsma, C.E.; Gul, N.; Koh, C.; Zhang, J.; Li, Y.M.; Maciaczyk, J.; et al. NOTCH pathway blockade depletes CD133-positive glioblastoma cells and inhibits growth of tumor neurospheres and xenografts. Stem. Cells 2010, 28, 5–16. [Google Scholar] [CrossRef] [Green Version]
  109. Wang, J.; Wakeman, T.P.; Lathia, J.D.; Hjelmeland, A.B.; Wang, X.F.; White, R.R.; Rich, J.N.; Sullenger, B.A. Notch promotes radioresistance of glioma stem cells. Stem. Cells 2010, 28, 17–28. [Google Scholar] [CrossRef] [Green Version]
  110. Xie, Q.; Wu, Q.; Kim, L.; Miller, T.E.; Liau, B.B.; Mack, S.C.; Yang, K.; Factor, D.C.; Fang, X.; Huang, Z.; et al. RBPJ maintains brain tumor-initiating cells through CDK9-mediated transcriptional elongation. J. Clin. Investig. 2016, 126, 2757–2772. [Google Scholar] [CrossRef]
  111. Natarajan, S.; Li, Y.; Miller, E.E.; Shih, D.J.; Taylor, M.D.; Stearns, T.M.; Bronson, R.T.; Ackerman, S.L.; Yoon, J.K.; Yun, K. Notch1-induced brain tumor models the sonic hedgehog subgroup of human medulloblastoma. Cancer Res. 2013, 73, 5381–5390. [Google Scholar] [CrossRef] [Green Version]
  112. Pierfelice, T.J.; Schreck, K.C.; Dang, L.; Asnaghi, L.; Gaiano, N.; Eberhart, C.G. Notch3 activation promotes invasive glioma formation in a tissue site-specific manner. Cancer Res. 2011, 71, 1115–1125. [Google Scholar] [CrossRef] [Green Version]
  113. Hovinga, K.E.; Shimizu, F.; Wang, R.; Panagiotakos, G.; Van Der Heijden, M.; Moayedpardazi, H.; Correia, A.S.; Soulet, D.; Major, T.; Menon, J.; et al. Inhibition of notch signaling in glioblastoma targets cancer stem cells via an endothelial cell intermediate. Stem. Cells 2010, 28, 1019–1029. [Google Scholar] [CrossRef] [Green Version]
  114. Wu, H.C.; Lin, Y.C.; Liu, C.H.; Chung, H.C.; Wang, Y.T.; Lin, Y.W.; Ma, H.I.; Tu, P.H.; Lawler, S.E.; Chen, R.H. USP11 regulates PML stability to control Notch-induced malignancy in brain tumours. Nat. Commun. 2014, 5, 3214. [Google Scholar] [CrossRef] [Green Version]
  115. Katsushima, K.; Natsume, A.; Ohka, F.; Shinjo, K.; Hatanaka, A.; Ichimura, N.; Sato, S.; Takahashi, S.; Kimura, H.; Totoki, Y.; et al. Targeting the Notch-regulated non-coding RNA TUG1 for glioma treatment. Nat. Commun. 2016, 7, 13616. [Google Scholar] [CrossRef] [PubMed]
  116. Sarkar, S.; Mirzaei, R.; Zemp, F.J.; Wei, W.; Senger, D.L.; Robbins, S.M.; Yong, V.W. Activation of NOTCH Signaling by Tenascin-C Promotes Growth of Human Brain Tumor-Initiating Cells. Cancer Res. 2017, 77, 3231–3243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Parmigiani, E.; Taylor, V.; Giachino, C. Tenascin-C induces expression of the Notch ligand Jagged1 to promote glioma growth. Transl. Cancer Res. 2017, 6, S1057–S1060. [Google Scholar] [CrossRef]
  118. Sivasankaran, B.; Degen, M.; Ghaffari, A.; Hegi, M.E.; Hamou, M.F.; Ionescu, M.C.; Zweifel, C.; Tolnay, M.; Wasner, M.; Mergenthaler, S.; et al. Tenascin-C is a novel RBPJkappa-induced target gene for Notch signaling in gliomas. Cancer Res. 2009, 69, 458–465. [Google Scholar] [CrossRef] [Green Version]
  119. Lim, K.J.; Brandt, W.D.; Heth, J.A.; Muraszko, K.M.; Fan, X.; Bar, E.E.; Eberhart, C.G. Lateral inhibition of Notch signaling in neoplastic cells. Oncotarget 2015, 6, 1666–1677. [Google Scholar] [CrossRef] [Green Version]
  120. Doetsch, F.; Caille, I.; Lim, D.A.; Garcia-Verdugo, J.M.; Alvarez-Buylla, A. Subventricular zone astrocytes are neural stem cells in the adult mammalian brain. Cell 1999, 97, 703–716. [Google Scholar] [CrossRef] [Green Version]
  121. Doetsch, F.; Garcia-Verdugo, J.M.; Alvarez-Buylla, A. Regeneration of a germinal layer in the adult mammalian brain. Proc. Natl. Acad. Sci. USA 1999, 96, 11619–11624. [Google Scholar] [CrossRef] [Green Version]
  122. Liau, B.B.; Sievers, C.; Donohue, L.K.; Gillespie, S.M.; Flavahan, W.A.; Miller, T.E.; Venteicher, A.S.; Hebert, C.H.; Carey, C.D.; Rodig, S.J.; et al. Adaptive Chromatin Remodeling Drives Glioblastoma Stem Cell Plasticity and Drug Tolerance. Cell Stem. Cell 2017, 20, 233–246.e7. [Google Scholar] [CrossRef] [Green Version]
  123. Eyler, C.E.; Matsunaga, H.; Hovestadt, V.; Vantine, S.J.; van Galen, P.; Bernstein, B.E. Single-cell lineage analysis reveals genetic and epigenetic interplay in glioblastoma drug resistance. Genome Biol. 2020, 21, 174. [Google Scholar] [CrossRef]
  124. Park, N.I.; Guilhamon, P.; Desai, K.; McAdam, R.F.; Langille, E.; O’Connor, M.; Lan, X.; Whetstone, H.; Coutinho, F.J.; Vanner, R.J.; et al. ASCL1 Reorganizes Chromatin to Direct Neuronal Fate and Suppress Tumorigenicity of Glioblastoma Stem Cells. Cell Stem. Cell 2017, 21, 411. [Google Scholar] [CrossRef]
  125. Rajakulendran, N.; Rowland, K.J.; Selvadurai, H.J.; Ahmadi, M.; Park, N.I.; Naumenko, S.; Dolma, S.; Ward, R.J.; So, M.; Lee, L.; et al. Wnt and Notch signaling govern self-renewal and differentiation in a subset of human glioblastoma stem cells. Genes Dev. 2019, 33, 498–510. [Google Scholar] [CrossRef]
  126. Shen, Q.; Wang, Y.; Kokovay, E.; Lin, G.; Chuang, S.M.; Goderie, S.K.; Roysam, B.; Temple, S. Adult SVZ stem cells lie in a vascular niche: A quantitative analysis of niche cell-cell interactions. Cell Stem. Cell 2008, 3, 289–300. [Google Scholar] [CrossRef] [Green Version]
  127. Tavazoie, M.; Van der Veken, L.; Silva-Vargas, V.; Louissaint, M.; Colonna, L.; Zaidi, B.; Garcia-Verdugo, J.M.; Doetsch, F. A specialized vascular niche for adult neural stem cells. Cell Stem. Cell 2008, 3, 279–288. [Google Scholar] [CrossRef] [Green Version]
  128. Ottone, C.; Krusche, B.; Whitby, A.; Clements, M.; Quadrato, G.; Pitulescu, M.E.; Adams, R.H.; Parrinello, S. Direct cell-cell contact with the vascular niche maintains quiescent neural stem cells. Nat. Cell Biol. 2014, 16, 1045–1056. [Google Scholar] [CrossRef] [Green Version]
  129. Calabrese, C.; Poppleton, H.; Kocak, M.; Hogg, T.L.; Fuller, C.; Hamner, B.; Oh, E.Y.; Gaber, M.W.; Finklestein, D.; Allen, M.; et al. A perivascular niche for brain tumor stem cells. Cancer Cell 2007, 11, 69–82. [Google Scholar] [CrossRef] [Green Version]
  130. Man, J.; Yu, X.; Huang, H.; Zhou, W.; Xiang, C.; Miele, L.; Liu, Z.; Bebek, G.; Bao, S.; Yu, J.S. Hypoxic Induction of Vasorin Regulates Notch1 Turnover to Maintain Glioma Stem-like Cells. Cell Stem. Cell 2018, 22, 104–118.e6. [Google Scholar] [CrossRef] [Green Version]
  131. Rusu, P.; Shao, C.; Neuerburg, A.; Acikgoz, A.A.; Wu, Y.; Zou, P.; Phapale, P.; Shankar, T.S.; Doring, K.; Dettling, S.; et al. GPDfile. Cell Stem. Cell 2019, 25, 241–257.e8. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, J.; Xu, S.L.; Duan, J.J.; Yi, L.; Guo, Y.F.; Shi, Y.; Li, L.; Yang, Z.Y.; Liao, X.M.; Cai, J.; et al. Invasion of white matter tracts by glioma stem cells is regulated by a NOTCH1-SOX2 positive-feedback loop. Nat. Neurosci. 2019, 22, 91–105. [Google Scholar] [CrossRef]
  133. Bettegowda, C.; Agrawal, N.; Jiao, Y.; Sausen, M.; Wood, L.D.; Hruban, R.H.; Rodriguez, F.J.; Cahill, D.P.; McLendon, R.; Riggins, G.; et al. Mutations in CIC and FUBP1 contribute to human oligodendroglioma. Science 2011, 333, 1453–1455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Halani, S.H.; Yousefi, S.; Velazquez Vega, J.; Rossi, M.R.; Zhao, Z.; Amrollahi, F.; Holder, C.A.; Baxter-Stoltzfus, A.; Eschbacher, J.; Griffith, B.; et al. Multi-faceted computational assessment of risk and progression in oligodendroglioma implicates NOTCH and PI3K pathways. Npj Precis Oncol. 2018, 2, 24. [Google Scholar] [CrossRef] [PubMed]
  135. Aoki, K.; Nakamura, H.; Suzuki, H.; Matsuo, K.; Kataoka, K.; Shimamura, T.; Motomura, K.; Ohka, F.; Shiina, S.; Yamamoto, T.; et al. Prognostic relevance of genetic alterations in diffuse lower-grade gliomas. Neuro Oncol. 2018, 20, 66–77. [Google Scholar] [CrossRef] [Green Version]
  136. Boulay, J.L.; Miserez, A.R.; Zweifel, C.; Sivasankaran, B.; Kana, V.; Ghaffari, A.; Luyken, C.; Sabel, M.; Zerrouqi, A.; Wasner, M.; et al. Loss of NOTCH2 positively predicts survival in subgroups of human glial brain tumors. PLoS ONE 2007, 2, e576. [Google Scholar] [CrossRef]
  137. Kondo, T. Molecular mechanisms involved in gliomagenesis. Brain Tumor. Pathol. 2017, 34, 1–7. [Google Scholar] [CrossRef]
  138. Ge, X.; Xiao, G.; Huang, H.; Du, J.; Tao, Y.; Yang, A.; Wu, H.; Zhang, Z.; Qiu, M. Stage-dependent regulation of oligodendrocyte development and enhancement of myelin repair by dominant negative Master-mind 1 protein. Glia 2019, 67, 1654–1666. [Google Scholar] [CrossRef]
  139. Galvao, R.P.; Kasina, A.; McNeill, R.S.; Harbin, J.E.; Foreman, O.; Verhaak, R.G.; Nishiyama, A.; Miller, C.R.; Zong, H. Transformation of quiescent adult oligodendrocyte precursor cells into malignant glioma through a multistep reactivation process. Proc. Natl. Acad. Sci. USA 2014, 111, E4214–E4223. [Google Scholar] [CrossRef] [Green Version]
  140. Ichimura, K.; Vogazianou, A.P.; Liu, L.; Pearson, D.M.; Backlund, L.M.; Plant, K.; Baird, K.; Langford, C.F.; Gregory, S.G.; Collins, V.P. 1p36 is a preferential target of chromosome 1 deletions in astrocytic tumours and homozygously deleted in a subset of glioblastomas. Oncogene 2008, 27, 2097–2108. [Google Scholar] [CrossRef] [Green Version]
  141. Muller, F.L.; Colla, S.; Aquilanti, E.; Manzo, V.E.; Genovese, G.; Lee, J.; Eisenson, D.; Narurkar, R.; Deng, P.; Nezi, L.; et al. Passenger deletions generate therapeutic vulnerabilities in cancer. Nature 2012, 488, 337–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Urban, N.; Blomfield, I.M.; Guillemot, F. Quiescence of Adult Mammalian Neural Stem Cells: A Highly Regulated Rest. Neuron 2019, 104, 834–848. [Google Scholar] [CrossRef]
  143. Somasundaram, K.; Reddy, S.P.; Vinnakota, K.; Britto, R.; Subbarayan, M.; Nambiar, S.; Hebbar, A.; Samuel, C.; Shetty, M.; Sreepathi, H.K.; et al. Upregulation of ASCL1 and inhibition of Notch signaling pathway characterize progressive astrocytoma. Oncogene 2005, 24, 7073–7083. [Google Scholar] [CrossRef] [Green Version]
  144. Narayanan, A.; Gagliardi, F.; Gallotti, A.L.; Mazzoleni, S.; Cominelli, M.; Fagnocchi, L.; Pala, M.; Piras, I.S.; Zordan, P.; Moretta, N.; et al. The proneural gene ASCL1 governs the transcriptional subgroup affiliation in glioblastoma stem cells by directly repressing the mesenchymal gene NDRG1. Cell Death Differ. 2019, 26, 1813–1831. [Google Scholar] [CrossRef] [Green Version]
  145. Rheinbay, E.; Suva, M.L.; Gillespie, S.M.; Wakimoto, H.; Patel, A.P.; Shahid, M.; Oksuz, O.; Rabkin, S.D.; Martuza, R.L.; Rivera, M.N.; et al. An aberrant transcription factor network essential for Wnt signaling and stem cell maintenance in glioblastoma. Cell Rep. 2013, 3, 1567–1579. [Google Scholar] [CrossRef] [Green Version]
  146. Vue, T.Y.; Kollipara, R.K.; Borromeo, M.D.; Smith, T.; Mashimo, T.; Burns, D.K.; Bachoo, R.M.; Johnson, J.E. ASCL1 regulates neurodevelopmental transcription factors and cell cycle genes in brain tumors of glioma mouse models. Glia 2020. [Google Scholar] [CrossRef]
  147. Barone, C.; Buccarelli, M.; Alessandrini, F.; Pagin, M.; Rigoldi, L.; Sambruni, I.; Favaro, R.; Ottolenghi, S.; Pallini, R.; Ricci-Vitiani, L.; et al. Sox2-dependent maintenance of mouse oligodendroglioma involves the Sox2-mediated downregulation of Cdkn2b, Ebf1, Zfp423, and Hey2. Glia 2020. [Google Scholar] [CrossRef] [PubMed]
  148. Havrda, M.C.; Paolella, B.R.; Ran, C.; Jering, K.S.; Wray, C.M.; Sullivan, J.M.; Nailor, A.; Hitoshi, Y.; Israel, M.A. Id2 mediates oligodendrocyte precursor cell maturation arrest and is tumorigenic in a PDGF-rich microenvironment. Cancer Res. 2014, 74, 1822–1832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Mehta, S.; Huillard, E.; Kesari, S.; Maire, C.L.; Golebiowski, D.; Harrington, E.P.; Alberta, J.A.; Kane, M.F.; Theisen, M.; Ligon, K.L.; et al. The central nervous system-restricted transcription factor Olig2 opposes p53 responses to genotoxic damage in neural progenitors and malignant glioma. Cancer Cell 2011, 19, 359–371. [Google Scholar] [CrossRef] [Green Version]
  150. Chow, R.D.; Guzman, C.D.; Wang, G.; Schmidt, F.; Youngblood, M.W.; Ye, L.; Errami, Y.; Dong, M.B.; Martinez, M.A.; Zhang, S.; et al. AAV-mediated direct in vivo CRISPR screen identifies functional suppressors in glioblastoma. Nat. Neurosci. 2017, 20, 1329–1341. [Google Scholar] [CrossRef] [Green Version]
  151. Andersson, E.R.; Lendahl, U. Therapeutic modulation of Notch signalling--are we there yet? Nat. Rev. Drug Discov. 2014, 13, 357–378. [Google Scholar] [CrossRef]
  152. Krop, I.; Demuth, T.; Guthrie, T.; Wen, P.Y.; Mason, W.P.; Chinnaiyan, P.; Butowski, N.; Groves, M.D.; Kesari, S.; Freedman, S.J.; et al. Phase I pharmacologic and pharmacodynamic study of the gamma secretase (Notch) inhibitor MK-0752 in adult patients with advanced solid tumors. J. Clin. Oncol. 2012, 30, 2307–2313. [Google Scholar] [CrossRef]
  153. Guner, G.; Lichtenthaler, S.F. The substrate repertoire of gamma-secretase/presenilin. Semin Cell Dev. Biol. 2020. [Google Scholar] [CrossRef]
  154. Lehal, R.; Zaric, J.; Vigolo, M.; Urech, C.; Frismantas, V.; Zangger, N.; Cao, L.; Berger, A.; Chicote, I.; Loubery, S.; et al. Pharmacological disruption of the Notch transcription factor complex. Proc. Natl. Acad. Sci. USA 2020, 117, 16292–16301. [Google Scholar] [CrossRef]
  155. 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. gamma-Secretase inhibitors in cancer clinical trials are pharmacologically and functionally distinct. Embo Mol. Med. 2017, 9, 950–966. [Google Scholar] [CrossRef]
  156. Saito, N.; Hirai, N.; Aoki, K.; Suzuki, R.; Fujita, S.; Nakayama, H.; Hayashi, M.; Ito, K.; Sakurai, T.; Iwabuchi, S. The Oncogene Addiction Switch from NOTCH to PI3K Requires Simultaneous Targeting of NOTCH and PI3K Pathway Inhibition in Glioblastoma. Cancers 2019, 11. [Google Scholar] [CrossRef] [Green Version]
  157. Zhang, C.; Martinez-Ledesma, E.; Gao, F.; Zhang, W.; Ding, J.; Wu, S.; Li, X.; Wu, J.; Yuan, Y.; Koul, D.; et al. Wild-type TP53 defined gamma-secretase inhibitor sensitivity and synergistic activity with doxorubicin in GSCs. Am. J. Cancer Res. 2019, 9, 1734–1745. [Google Scholar]
  158. Fassl, A.; Tagscherer, K.E.; Richter, J.; Berriel Diaz, M.; Alcantara Llaguno, S.R.; Campos, B.; Kopitz, J.; Herold-Mende, C.; Herzig, S.; Schmidt, M.H.; et al. Notch1 signaling promotes survival of glioblastoma cells via EGFR-mediated induction of anti-apoptotic Mcl-1. Oncogene 2012, 31, 4698–4708. [Google Scholar] [CrossRef]
  159. Yin, J.; Park, G.; Kim, T.H.; Hong, J.H.; Kim, Y.J.; Jin, X.; Kang, S.; Jung, J.E.; Kim, J.Y.; Yun, H.; et al. Pigment Epithelium-Derived Factor (PEDF) Expression Induced by EGFRvIII Promotes Self-renewal and Tumor Progression of Glioma Stem Cells. PLoS Biol. 2015, 13, e1002152. [Google Scholar] [CrossRef] [Green Version]
  160. Purow, B.W.; Sundaresan, T.K.; Burdick, M.J.; Kefas, B.A.; Comeau, L.D.; Hawkinson, M.P.; Su, Q.; Kotliarov, Y.; Lee, J.; Zhang, W.; et al. Notch-1 regulates transcription of the epidermal growth factor receptor through p53. Carcinogenesis 2008, 29, 918–925. [Google Scholar] [CrossRef] [Green Version]
  161. Verhaak, R.G.; Hoadley, K.A.; Purdom, E.; Wang, V.; Qi, Y.; Wilkerson, M.D.; Miller, C.R.; Ding, L.; Golub, T.; Mesirov, J.P.; et al. Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell 2010, 17, 98–110. [Google Scholar] [CrossRef] [Green Version]
  162. Lu, F.; Chen, Y.; Zhao, C.; Wang, H.; He, D.; Xu, L.; Wang, J.; He, X.; Deng, Y.; Lu, E.E.; et al. Olig2-Dependent Reciprocal Shift in PDGF and EGF Receptor Signaling Regulates Tumor Phenotype and Mitotic Growth in Malignant Glioma. Cancer Cell 2016, 29, 669–683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Alcolea, M.P.; Greulich, P.; Wabik, A.; Frede, J.; Simons, B.D.; Jones, P.H. Differentiation imbalance in single oesophageal progenitor cells causes clonal immortalization and field change. Nat. Cell Biol. 2014, 16, 615–622. [Google Scholar] [CrossRef] [Green Version]
  164. Bowling, S.; Lawlor, K.; Rodriguez, T.A. Cell competition: The winners and losers of fitness selection. Development 2019, 146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Friebel, E.; Kapolou, K.; Unger, S.; Nunez, N.G.; Utz, S.; Rushing, E.J.; Regli, L.; Weller, M.; Greter, M.; Tugues, S.; et al. Single-Cell Mapping of Human Brain Cancer Reveals Tumor-Specific Instruction of Tissue-Invading Leukocytes. Cell 2020, 181, 1626–1642.e20. [Google Scholar] [CrossRef]
  166. Klemm, F.; Maas, R.R.; Bowman, R.L.; Kornete, M.; Soukup, K.; Nassiri, S.; Brouland, J.P.; Iacobuzio-Donahue, C.A.; Brennan, C.; Tabar, V.; et al. Interrogation of the Microenvironmental Landscape in Brain Tumors Reveals Disease-Specific Alterations of Immune Cells. Cell 2020, 181, 1643–1660.e17. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The NOTCH pathway can act either as an oncogene or as a tumor suppressor in glioma, depending on the context. On the one hand, NOTCH signaling activity in subpopulations of glioma cells can enhance stem cell features, promote resistance to radio- and chemo-therapies, and favor tumor development by activating oncogenic pathways (e.g., PI3K/Akt) or inhibiting tumor suppressors (e.g., PML). NOTCH can also regulate long non-coding RNAs such as TUG1 to maintain stemness and suppress differentiation. Moreover, NOTCH signaling can establish positive feedback loops with Tenascin-C, EGFR, and a SOX9-SOX2 axis. Finally, NOTCH can modulate interactions between glioma stem cells (GSCs) and their niche in different locations within the tumor mass, including in hypoxic regions (through Vasorin) and at the invasive front (through SOX9-SOX2), thereby promoting stem cell features and invasive potential. On the other hand, NOTCH-inactivating mutations and low expression levels of canonical NOTCH target genes have been identified in patients with glioma subtypes, and genetic NOTCH inhibition accelerates glioma formation in glioma mouse models, pointing to a tumor-suppressive role of NOTCH in glioma similar to in epithelial cancers. Although the molecular mechanisms driving a tumor-suppressive role of NOTCH signaling in glioma are still largely unknown, NOTCH can inhibit tumor cell proliferation and glioma growth by suppressing ASCL1 and OLIG2 expression. Interestingly, NOTCH-mediated suppression of ASCL1 can result in either oncogenic or tumor-suppressive effects by inhibiting differentiation or proliferation of glioma cells, respectively.
Figure 1. The NOTCH pathway can act either as an oncogene or as a tumor suppressor in glioma, depending on the context. On the one hand, NOTCH signaling activity in subpopulations of glioma cells can enhance stem cell features, promote resistance to radio- and chemo-therapies, and favor tumor development by activating oncogenic pathways (e.g., PI3K/Akt) or inhibiting tumor suppressors (e.g., PML). NOTCH can also regulate long non-coding RNAs such as TUG1 to maintain stemness and suppress differentiation. Moreover, NOTCH signaling can establish positive feedback loops with Tenascin-C, EGFR, and a SOX9-SOX2 axis. Finally, NOTCH can modulate interactions between glioma stem cells (GSCs) and their niche in different locations within the tumor mass, including in hypoxic regions (through Vasorin) and at the invasive front (through SOX9-SOX2), thereby promoting stem cell features and invasive potential. On the other hand, NOTCH-inactivating mutations and low expression levels of canonical NOTCH target genes have been identified in patients with glioma subtypes, and genetic NOTCH inhibition accelerates glioma formation in glioma mouse models, pointing to a tumor-suppressive role of NOTCH in glioma similar to in epithelial cancers. Although the molecular mechanisms driving a tumor-suppressive role of NOTCH signaling in glioma are still largely unknown, NOTCH can inhibit tumor cell proliferation and glioma growth by suppressing ASCL1 and OLIG2 expression. Interestingly, NOTCH-mediated suppression of ASCL1 can result in either oncogenic or tumor-suppressive effects by inhibiting differentiation or proliferation of glioma cells, respectively.
Cells 09 02304 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Parmigiani, E.; Taylor, V.; Giachino, C. Oncogenic and Tumor-Suppressive Functions of NOTCH Signaling in Glioma. Cells 2020, 9, 2304. https://doi.org/10.3390/cells9102304

AMA Style

Parmigiani E, Taylor V, Giachino C. Oncogenic and Tumor-Suppressive Functions of NOTCH Signaling in Glioma. Cells. 2020; 9(10):2304. https://doi.org/10.3390/cells9102304

Chicago/Turabian Style

Parmigiani, Elena, Verdon Taylor, and Claudio Giachino. 2020. "Oncogenic and Tumor-Suppressive Functions of NOTCH Signaling in Glioma" Cells 9, no. 10: 2304. https://doi.org/10.3390/cells9102304

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