Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3161 2022-05-10 16:02:14 |
2 Please edit title of entry to "YAP/TAZ Activation in Head and Neck Cancer" Meta information modification 3161 2022-05-10 17:13:18 | |
3 format corrected. -6 word(s) 3155 2022-05-11 03:43:56 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Faraji, F.; , .; Gutkind, J. YAP/TAZ Activation in Head and Neck Cancer. Encyclopedia. Available online: https://encyclopedia.pub/entry/22774 (accessed on 24 April 2024).
Faraji F,  , Gutkind J. YAP/TAZ Activation in Head and Neck Cancer. Encyclopedia. Available at: https://encyclopedia.pub/entry/22774. Accessed April 24, 2024.
Faraji, Farhoud, , J.silvio Gutkind. "YAP/TAZ Activation in Head and Neck Cancer" Encyclopedia, https://encyclopedia.pub/entry/22774 (accessed April 24, 2024).
Faraji, F., , ., & Gutkind, J. (2022, May 10). YAP/TAZ Activation in Head and Neck Cancer. In Encyclopedia. https://encyclopedia.pub/entry/22774
Faraji, Farhoud, et al. "YAP/TAZ Activation in Head and Neck Cancer." Encyclopedia. Web. 10 May, 2022.
YAP/TAZ Activation in Head and Neck Cancer
Edit

The Hippo signaling pathway, originally discovered as a mechanism regulating tissue growth and organ size, transduces intracellular and extracellular signals to regulate the transcriptional co-activators YAP and TAZ. Alterations in the Hippo pathway resulting in persistent YAP and TAZ activation have emerged as major oncogenic drivers. The researchers' analysis of the human Head and neck squamous cell carcinoma (HNSCC) oncogenome revealed multiple genomic alterations impairing Hippo signaling and activating YAP and TAZ, which in turn contribute to HNSCC development. This includes mutations and deletions of the FAT1 gene (29%) and amplification of the WWTR1 (encoding TAZ, 14%) and YAP1 genes (8%), together representing one of the most genetically altered signaling mechanisms in this malignancy. 

head and neck squamous cell carcinoma HNSC HNSCC Hippo YAP TAZ FAT1

1. Mechanisms of YAP/TAZ Activation in Head and Neck Squamous Cell Carcinoma (HNSCC)

1.1. FAT1: A Membrane-Associated Proto-Cadherin Assembling the Hippo Signalome

Although dysregulation of the Hippo pathway occurs frequently in multiple human malignancies, alterations in the canonical core Hippo kinases are rarely observed. In contrast, with an alteration rate of nearly 30%, FAT1 mutation constitutes one of the most frequent mutations in HNSCC [1]. Inactivating FAT1 mutations, primarily in the form of homozygous deletions, but also in the form of protein-truncating nonsense mutations, strongly suggest that FAT1 functions as a tumor suppressor [1][2][3].
Under physiological conditions, FAT1 serves as a membrane-associated scaffold to organize the activated core Hippo signaling complex (signalome). Through loss of function, ectopic expression, and protein interaction experiments, TAOK1/2/3 were demonstrated to represent the upstream kinases triggered by FAT1 to activate the core Hippo complex [1]. In this context, the FAT1 intracellular domain served to assemble the multimeric Hippo core signalome consisting of phospho-MST1 (p-MST1), p-MST2, p-SAV1, p-LATS1, and p-MOB1, culminating in YAP phosphorylation. The degree to which Hippo-mediated YAP phosphorylation is dependent upon FAT1 is underscored by the functional loss of FAT1, which results in the disassembly of the Hippo signalome with consequent YAP activation, nuclear localization, and activation of YAP-mediated proliferative, survival, and tumorigenic transcriptional programs. These mechanistic observations were validated using cell-based functional assays that demonstrated FAT1 expression abrogated YAP-activated transcription and cell proliferation, which was reversed upon co-expression of a Hippo-insensitive YAP mutant [1]. This research was the first to mechanistically link FAT1 disruption with Hippo signaling disruption and uncontrolled YAP activation. Interestingly, FAT1 mutation has been identified in carcinogen-induced models of HNSCC, potentially linking carcinogen-induced mutation and Hippo pathway perturbation as mediators of HPV-negative HNSCC initiation [4]. FAT2, FAT3, and FAT4 are also recurrently altered in human HNSCC and in carcinogen-induced murine models of HNSCC [4][5]. However, the biological significance of these alterations remains to be examined [6]. While inactivating FAT1 mutations are uncommon in HPV-positive HNSCC [7], preliminary analyses suggest that Hippo pathway perturbation and YAP/TAZ activation are also widespread in HPV-positive HNSCC (not shown), thus representing an area of active investigation.

1.2. FAT1-Independent Mechanisms of Oncogenic Hippo Pathway Perturbation in HNSCC

Independent of FAT1 mutation, YAP activation is a prevalent feature of HNSCC. Epidermal growth factor receptor (EGFR) activation also represents a common event in HNSCC that is correlated with aggressive disease [8][9]. Recent studies have described a series of mechanisms through which EGFR and receptor tyrosine kinase downstream signaling cascades via RAS-MAPK and phosphoinositide-3-kinase (PI3K) can drive oncogenic YAP activation.
Amplification and gain of function mutations in PIK3CA result in the hyperactivation of its gene product PI3K and PI3K signaling pathway activation in HNSCC. PI3K signaling has been shown to induce nuclear YAP localization and YAP-target transcription in response to EGFR or GPCR agonists. Demonstrating strict signaling through PI3K and PDK1, this response was attenuated by PI3K and PDK1 inhibitors and siRNA mediated knockdown. Mechanistically, in cell contact-inhibited conditions, PDK1 binds to MST1 and LATS1 through SAV1, enabling Hippo activation and resulting in YAP phosphorylation and cytoplasmic retention, and cellular growth arrest. In the presence of growth factors or constitutive PI3K signaling, PDK1 is recruited to the plasma membrane, resulting in the dissociation of the MST1:SAV1:LATS1 complex, YAP nuclear accumulation, and cell cycle entry [10].
More recently, EGFR has been shown to activate YAP/TAZ independently of PI3K through the direct phosphorylation of MOB1A in HNSCC cells [11]. EGFR phosphorylates novel MOB1A tyrosine residues in the presence of epidermal growth factor, resulting in reductions in p-LATS1/2 and p-YAP, triggering nuclear translocation of YAP and activation of YAP-mediated transcription programs. Furthermore, inhibition of EGFR in HNSCC cells with erlotinib was shown to be sufficient in abrogating YAP activation and the expression of YAP-mediated transcriptional programs. While further studies are required to delineate the mechanism by which EGFR-mediated MOB1A phosphorylation prevents LATS1 activation, this research unveiled a novel, therapeutically actionable regulatory signaling node by which multiple receptors and non-receptor tyrosine kinases may converge with the Hippo pathway to control YAP/TAZ activity [11].
Disruption in core Hippo signaling has been shown to induce HNSCC in model systems. Genetic deletion of Mob1a and Mob1b in murine lingual epithelia led to rapid squamous cell carcinoma. Consistent with the canonical Hippo core signaling model, knockout of Mob1a/b in tongue epithelia resulted in diminished LATS1 but not MST1 protein, increased nuclear YAP, and upregulation in YAP target genes [12]. Much remains to be understood about core Hippo signaling and the functions of MOB1 in YAP/TAZ regulation. As illustrated in the setting of Mob1a/Mob1b knockout in tongue epithelia, knockout of Yap prevented carcinogenesis induced by Mob1a/b deletion, but knockout of Taz in the same context resulted in more aggressive SCC with more invading lesions [12]. While conditional deletion of Mob1a/b suggests that disruption of these core Hippo complex proteins quickly leads to carcinogenesis, Mob1a/b can also contribute to signaling cascades beyond Hippo. For example, Mob1a and Mob1b stimulate the activity of non-Hippo kinases including STK38 (NDR1) and STK28L (NDR2) [13]. While NDR1/2 have been implicated in Hippo signaling and reported to phosphorylate YAP [14], NDR1 also serves to attenuate mitogen activated kinase signaling as a negative regulator of MAP3K1/2 [15]. Indeed, collectively NDR1/2 exert pleiotropic potentially proto-oncogenic influences on cell cycle progression, apoptosis, stress signaling, and autophagy [16]. Exploring the complex interactions across Hippo and other signaling cascades may clarify seemingly paradoxical effects observed upon the genetic depletion of kinase-regulatory scaffolds, such as Mob1a/b.

2. Consequences YAP/TAZ Activation in HNSCC

2.1. Cancer Stemness

Chromosome 3q25-26 is a frequently amplified locus in HNSCC that harbors multiple cancer-associated genes including WWTR1 (TAZ), PIK3CA, and SOX2. Co-occurrence of 3q25-26 amplification and TP53 mutation is associated with poor prognosis in HNSCC [17]. Interestingly, a TAZ:TEAD4 complex has been shown to co-activate SOX2 transcription in HNSCC. In this context, SOX2 expression was found sufficient to promote cancer stem cell marker expression, tumor cell self-renewal in vitro, and tumor growth in vivo [18]. Importantly, knockdown of TAZ diminished these phenotypes and exogenous expression of SOX2 upon TAZ knockdown was sufficient to rescue them. Together, these findings show that TAZ- and TEAD4-mediated coactivation of SOX2 expression is sufficient to induce stemness programs in HNSCC and offer the possibility that TAZ activation may mediate transcriptional programs that lead to the poor prognosis associated with 3q25-26 copy gain. In line with these findings, an independent research found that TAZ activation promotes migration, invasion, and survival by an epithelial to mesenchymal transition-like program in HNSCC cells, which in turn promotes cancer stem cell maintenance and expansion [19].
The recurrently amplified segment of chromosome 3q also harbors two genes that are frequently co-amplified in HNSCC: TP63, a key regulator of epidermal cell differentiation and proliferation [20], and ACTL6A, a subunit of the SWI/SNF ATP-dependent chromatin remodeling complexes [21]. Interestingly, p63 forms a complex with ACTL6A and other SWI/SNF subunits in HNSCC to control a stem-like transcriptional program that enhances the regenerative potential of HNSCC cells in vitro, and promotes a pro-tumorigenic proliferative, undifferentiated cancer cell state in vivo. Importantly, p63 and ACTL6A were found to directly repress transcription of WWC1 [22]. The gene product of WWC1, Kibra, is a tumor suppressor that promotes Hippo pathway activation [23][24], thereby releasing Hippo-mediated inhibition of YAP/TAZ. Indeed, depletion of p63 or ACTL6A, increased Kibra expression, increased p-YAP, depressed nuclear translocation of YAP, and diminished HNSCC regenerative potential in vitro [22].

2.2. Tumor Progression and Poor Prognosis

Progressively increased YAP activation is a feature of oncogenic progression in diverse malignancies [25]. In the squamous epithelia of the upper aerodigestive tract, physiologically nuclear YAP is restricted to the basal epidermis [1][26]. The extent and degree of nuclear YAP have been shown to progressively increase with worsening histological severity in oral premalignant lesions and culminate with diffuse, strong nuclear YAP expression in the majority of cells in HNSCC. Furthermore, the frequency of nuclear YAP-positive tumor cells also increases with worsening histological grade in HNSCC [1][27]. Array-based transcriptome profiling upon YAP/TAZ knockdown in HNSCC cell lines has elucidated YAP/TAZ-mediated transcription programs in HNSCC. The resultant YAP/TAZ-regulated gene signature, defined as the set of transcripts downregulated upon YAP/TAZ knockdown, showed enrichment for stemness, self-renewal, cell cycle progression, and invasion programs [27]. Consistent with these findings, it has also been shown that nuclear YAP and TAZ in HNSCC are enriched at the tumor invasive front [28][29]. In addition, increased nuclear YAP expression in primary tumor specimens has been associated with lymph node metastasis, further suggesting that YAP may play a role in driving invasion and metastatic programs [29]. In line with these findings, TAZ activation was also shown to be a prevalent feature in HNSCC cell lines and an independent prognostic factor for disease-free and overall survival in patients with tongue HNSCC [30].
Whether cancer associated mutations modify prognosis or therapeutic sensitivity are key questions that may shed insight into mechanisms of tumor initiation and progression, and aid in the identification of prognostic, diagnostic, and therapeutic biomarkers. In this regard, the potential roles of YAP/TAZ on prognosis and other clinicopathologic features in HNSCC have been evaluated. YAP copy number alteration (CNA) is observed in 8% of HNSCC. While YAP amplification is associated with increased YAP transcript abundance, augmented YAP expression is often observed in HNSCC in the absence of CNA [31]. Post-translational signaling cascades regulate YAP activation; therefore, YAP activity is likely not directly related to its transcriptional expression levels. Eun and colleagues interrogated the Cancer Genome Atlas (TCGA) database to identify transcripts correlated with YAP transcript expression and CNA in order to develop a gene expression signature for YAP activation [31]. This YAP activation signature was found to stratify prognosis with regard to disease-free survival, disease-specific survival, and overall survival in four independent patient cohorts. Importantly, the YAP activation signature stratified survival after multivariable correction for confounding clinical factors known to be associated with survival (age, tumor classification, nodal classification, TNM staging group, and anatomic site) [31].
Analyses of clinical data-linked, genome-wide databases have identified interactions between PIK3CA and YAP activation in HNSCC and demonstrated that these interactions may be associated with survival outcomes. While neither mutation nor CNA in PIK3CA was associated with recurrence-free survival (RFS), PIK3CA mRNA expression was found to be associated with RFS in TCGA and an independent dataset on single and multivariable analyses [32]. Consistent with mechanistic models in which PIK3CA activation induces downstream YAP dephosphorylation and nuclear translocation to drive YAP-mediated transcription programs, HNSCC tumor samples with high PIK3CA expression displayed lower abundances of p-YAP and PIK3CAhigh tumors expressed transcriptional programs enriched for YAP/TAZ target genes [32]. These provide translational relevance to the mechanistic studies linking PIK3CA activity to YAP/TAZ activation, suggest that the PIK3CA-YAP axis may drive aggressive forms of HNSCC, and provide a strong rationale to target YAP in PIK3CAhigh HNSCC.

2.3. Therapeutic Resistance

Combined with radiotherapy and surgical resection, cisplatin-based chemotherapy is a mainstay of curative-intent therapy for HNSCC [33]. Limited evidence suggests that short interfering RNA-mediated YAP-knockdown re-sensitizes cisplatin-resistant HNSCC cell lines to cisplatin. These findings suggest that YAP may be considered a potential therapeutic target for cisplatin-resistant HNSCC [34]. Using a similar line of experimentation, an independent group identified a ribosome-binding protein that mediates the interaction between the ribosome and the endoplasmic reticulum membrane, RRBP1, as a mediator of cisplatin resistance in HNSCC by augmenting YAP expression. In this setting, the use of a putative RRBP1 inhibitor, Radelozid, diminished YAP expression and sensitized HNSCC cells to cisplatin in both in vitro and in vivo assays [35]. In addition, TAZ overexpression has been shown to enhance resistance while YAP or TAZ knockdown sensitizes HNSCC cell lines to cisplatin and fluorouracil [19][36].
More recently, precision therapeutic targeting of the MAPK pathway, which is commonly activated in HNSCC and is associated with more aggressive tumor growth, nodal metastasis, and recurrence, has been evaluated in clinical trials [37][38][39]. In a window-of-opportunity trial in previously untreated HNSCC patients, the small molecule MEK inhibitor trametinib was shown to inhibit the MAPK pathway in 33% of patients, as evaluated by tumor p-ERK abundance, and to result in clinical tumor response rates in 65% of patients [40]. Given that single-modality targeting of MAPK signaling has not shown durable efficacy in a wide variety of tumor types [41], follow up studies have explored mechanisms of trametinib resistance in HNSCC that could lead to multimodal precision therapeutics with durable efficacy. In this regard, trametinib-resistant HNSCC cell lines were generated by culturing cells in sequentially increasing doses of trametinib. The resultant cells demonstrated at least 105-fold increased resistance to trametinib and elevated YAP activity as measured by transcriptional reporter assays and upregulation of canonical YAP-target genes. In addition, trametinib-treated patient-derived xenografts that escaped growth inhibition were found to have 15-fold greater abundances of unphosphorylated YAP protein and upregulated YAP-target genes. Accordingly, combining trametinib with verteporfin, which inhibits YAP signaling, showed synergistic effects. These results highlight the potential importance of YAP activation in mediating resistance to MAPK pathway inhibition and suggest YAP inhibition as a potential strategy to enhance the efficacy of MAPK pathway inhibition in patients with HNSCC [42].

3. YAP/TAZ Activation Exposes HNSCC Vulnerabilities and Therapeutic Opportunities

Accumulating evidence demonstrating YAP and TAZ as oncogenic effectors and essential cancer dependencies in HNSCC supports the development of novel oncotherapeutics against YAP/TAZ [43][44]. However, therapeutic targeting of transcription factors remains a major challenge. Beyond YAP and TAZ, numerous cancer-associated transcription factors have emerged as promising therapeutic targets [45]. Yet the development of specific small molecule therapeutics against transcription factors has been hampered by hurdles related to the transcription factors not typically possessing enzymatic activities and practical challenges of disrupting protein:protein and protSein:nucleic acid interactions [45].
The requirement for YAP/TAZ to interact with TEAD family DNA binding proteins present an opportunity for its inhibition. An in vitro functional screen in mammalian cells for small molecule inhibitors of YAP:TEAD based transcriptional transactivation identified three porphyrin derivatives protoporphyrin IX, hematoporphyrin, and verteporfin as inhibitors of YAP-induced transcription [46]. Coincidentally, an independent screen identified these same three porphyrins as disruptors of YAP:TEAD interaction in Drosophila. Further studies validated verteporfin as a strong inhibitor of YAP:TEAD interaction [46]. Verteporfin is an FDA-approved photosensitizer used in photodynamic therapy for the treatment of age-related macular degeneration, suggesting verteporfin exerts cellular effects beyond the inhibition of YAP:TEAD interactions (Figure 1) [47]. Despite being studied for the past decade as a YAP inhibitor, structural mechanistic details about its interaction with YAP/TAZ remain unclear, and its effects on cellular targets other than YAP/TAZ, including autophagosome inhibition, suggesting that verteporfin constitutes a nonspecific YAP/TAZ inhibitor [48].
Figure 1. Targeting Hippo and YAP/TAZ. Therapeutic agents promoting Hippo activation or YAP/TAZ inhibition are shown in red. Grey arrows and molecules indicate inhibition of downstream processes. Cetuximab-mediated EGFR inhibition, Alpelisib-mediated PI3K inhibition, Dasatinib-mediated SRC inhibition, Defactinib and IN10018-mediated FAK inhibition, and Trametinib-mediated MEK inhibition relieve Hippo-inhibitory inputs and enable Hippo signaling to inhibit YAP/TAZ. Tankyrase inhibitors prevent proteolytic degradation of AMOT, enabling AMOT to activate Hippo signaling. Small molecule TEAD inhibitors (smTEADi) prevent YAP/TAZ:TEAD interaction by inhibiting TEAD autopalmitoylation. Verteporfin inhibits YAP/TAZ:TEAD interaction [40][46][49][50][51][52][53][54][55].
Further research of the YAP/TAZ:TEAD interaction has revealed a unique mechanism through which YAP/TAZ and TEAD family proteins interact to co-activate transcription and has unveiled a novel opportunity for pharmacologic inhibition. YAP/TAZ and TEAD complex formation require an autopalmitoylation event in which TEAD catalyzes the covalent attachment of a palmitate fatty acid to itself. TEAD autopalmitoylation is both necessary for YAP/TAZ:TEAD complex formation and YAP/TAZ-mediated transcriptional co-activation [56]. Small molecule inhibitors of TEAD autopalmitoylation have recently been demonstrated to bind TEAD proteins and prevent TEAD autopalmitoylation, disrupt YAP/TAZ interaction with TEAD proteins, diminish YAP/TAZ:TEAD transcriptional co-activation, downregulate YAP/TAZ:TEAD target transcripts, and inhibit in vivo tumor growth of NF2-deficient cancer cell lines (Figure 1) [50]. Whether TEAD autopalmitoylation inhibitors show activity against YAP-activated HNSCC remains an open question and an area of active investigation.
Numerous cellular mechanisms regulate YAP/TAZ activity. Thus, YAP/TAZ inhibition could be achieved through a variety of therapeutic avenues beyond the disruption of YAP/TAZ:TEAD interactions. The GPCR-PKA signaling axis modulates Hippo-YAP/TAZ activity. GPCRs represent one of the most frequently targeted protein classes, with small molecule and peptide drugs designed to target essentially every type of GPCR either approved for clinical use or in clinical development [57][58]. Pharmacological inhibition of G12/13 or Gq/11, or pharmacological activation of Gs could achieve downstream YAP/TAZ inhibition. Attempting to target the Hippo pathway far upstream of YAP/TAZ at plasma membrane bound receptors, however, could result in unforeseen effects [59][60].
PKA may represent another therapeutic target in GPCR-PKA-mediated regulation of the Hippo pathway. Interestingly, PKA inhibition via phosphatase PP2A activation has shown antineoplastic activity against models of small cell carcinoma [61]. Given that PP2A drives YAP/TAZ activation, the therapeutic use of PP2A phosphatase activators is unlikely to be beneficial in the treatment of HNSCC. However, drugs that target and inactivate PP2A may prove beneficial in the treatment of HNSCC. PP2A inhibitors are in current development and have shown efficacy against a variety of YAP-activated tumor types [62][63].
The interaction of β-catenin and YAP/TAZ also presents a potential therapeutically actionable opportunity. Wnt/β-catenin signaling has independently been associated with tumor initiation and progression and, as a result, multiple inhibitors of this pathway have been developed [64]. Given its role in downregulating β-catenin activity, the β-catenin destruction complex has been a focus for therapeutic intervention with the development of tankyrase inhibitors [60][61]. Tankyrase (TNKS) interacts with and degrades the β-catenin destruction complex component Axin via ubiquitin-mediated proteasomal degradation [65][66]. Tankyrase inhibitors and other compounds that stabilize Axin have been shown to inhibit both YAP/TAZ and β-catenin activity through the cytoplasmic retention of YAP/TAZ and subsequent restoration of destruction complex activity (Figure 1) [44][67]. In this manner, tankyrase inhibitors promote YAP/TAZ inhibition by the β-catenin destruction complex. More recently, tankyrase inhibitors have been shown to also exert a more direct effect on the Hippo pathway by stabilizing AMOT family proteins, which promote the cytoplasmic retention of YAP/TAZ and prevent YAP/TAZ induced transcriptional programs [51][68].

References

  1. Martin, D.; Degese, M.S.; Vitale-Cross, L.; Iglesias-Bartolome, R.; Valera, J.L.C.; Wang, Z.; Feng, X.; Yeerna, H.; Vadmal, V.; Moroishi, T.; et al. Assembly and activation of the Hippo signalome by FAT1 tumor suppressor. Nat. Commun. 2018, 9, 2372.
  2. Morris, L.G.T.; Kaufman, A.M.; Gong, Y.; Ramaswami, D.; Walsh, L.A.; Turcan, Ş.; Eng, S.; Kannan, K.; Zou, Y.; Peng, L.; et al. Recurrent somatic mutation of FAT1 in multiple human cancers leads to aberrant Wnt activation. Nat. Genet. 2013, 45, 253–261.
  3. Pastushenko, I.; Mauri, F.; Song, Y.; de Cock, F.; Meeusen, B.; Swedlund, B.; Impens, F.; Van Haver, D.; Opitz, M.; Thery, M.; et al. Fat1 deletion promotes hybrid EMT state, tumour stemness and metastasis. Nature 2021, 589, 448–455.
  4. Wang, Z.; Wu, V.H.; Allevato, M.M.; Gilardi, M.; He, Y.; Luis Callejas-Valera, J.; Vitale-Cross, L.; Martin, D.; Amornphimoltham, P.; Mcdermott, J.; et al. Syngeneic animal models of tobacco-associated oral cancer reveal the activity of in situ anti-CTLA-4. Nat. Commun. 2019, 10, 5546.
  5. Mann, J.E.; Kulkarni, A.; Birkeland, A.C.; Kafelghazal, J.; Eisenberg, J.; Jewell, B.M.; Ludwig, M.L.; Spector, M.E.; Jiang, H.; Carey, T.E.; et al. The molecular landscape of the University of Michigan laryngeal squamous cell carcinoma cell line panel. Head Neck 2019, 41, 3114–3124.
  6. Katoh, M. Function and cancer genomics of FAT family genes (review). Int. J. Oncol. 2012, 41, 1913–1918.
  7. Kim, K.T.; Kim, B.-S.; Kim, J.H. Association between FAT1 mutation and overall survival in patients with human papillomavirus-negative head and neck squamous cell carcinoma. Head Neck 2016, 38 (Suppl. 1), E2021–E2029.
  8. Kriegs, M.; Clauditz, T.S.; Hoffer, K.; Bartels, J.; Buhs, S.; Gerull, H.; Zech, H.B.; Bußmann, L.; Struve, N.; Rieckmann, T.; et al. Analyzing expression and phosphorylation of the EGF receptor in HNSCC. Sci. Rep. 2019, 9, 13564.
  9. Temam, S.; Kawaguchi, H.; El-Naggar, A.K.; Jelinek, J.; Tang, H.; Liu, D.D.; Lang, W.; Issa, J.-P.; Lee, J.J.; Mao, L. Epidermal growth factor receptor copy number alterations correlate with poor clinical outcome in patients with head and neck squamous cancer. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2007, 25, 2164–2170.
  10. Fan, R.; Kim, N.-G.; Gumbiner, B.M. Regulation of Hippo pathway by mitogenic growth factors via phosphoinositide 3-kinase and phosphoinositide-dependent kinase-1. Proc. Natl. Acad. Sci. USA 2013, 110, 2569–2574.
  11. Ando, T.; Arang, N.; Wang, Z.; Costea, D.E.; Feng, X.; Goto, Y.; Izumi, H.; Gilardi, M.; Ando, K.; Gutkind, J.S. EGFR Regulates the Hippo pathway by promoting the tyrosine phosphorylation of MOB1. Commun. Biol. 2021, 4, 1237.
  12. Omori, H.; Nishio, M.; Masuda, M.; Miyachi, Y.; Ueda, F.; Nakano, T.; Sato, K.; Mimori, K.; Taguchi, K.; Hikasa, H.; et al. YAP1 is a potent driver of the onset and progression of oral squamous cell carcinoma. Sci. Adv. 2020, 6, eaay3324.
  13. Bichsel, S.J.; Tamaskovic, R.; Stegert, M.R.; Hemmings, B.A. Mechanism of activation of NDR (nuclear Dbf2-related) protein kinase by the hMOB1 protein. J. Biol. Chem. 2004, 279, 35228–35235.
  14. Zhang, L.; Tang, F.; Terracciano, L.; Hynx, D.; Kohler, R.; Bichet, S.; Hess, D.; Cron, P.; Hemmings, B.A.; Hergovich, A.; et al. NDR functions as a physiological YAP1 kinase in the intestinal epithelium. Curr. Biol. CB 2015, 25, 296–305.
  15. Enomoto, A.; Kido, N.; Ito, M.; Morita, A.; Matsumoto, Y.; Takamatsu, N.; Hosoi, Y.; Miyagawa, K. Negative regulation of MEKK1/2 signaling by serine-threonine kinase 38 (STK38). Oncogene 2008, 27, 1930–1938.
  16. Hergovich, A. The Roles of NDR Protein Kinases in Hippo Signalling. Genes 2016, 7, 21.
  17. Cheng, H.; Yang, X.; Si, H.; Saleh, A.D.; Xiao, W.; Coupar, J.; Gollin, S.M.; Ferris, R.L.; Issaeva, N.; Yarbrough, W.G.; et al. Genomic and Transcriptomic Characterization Links Cell Lines with Aggressive Head and Neck Cancers. Cell Rep. 2018, 25, 1332–1345.e5.
  18. Li, J.; Li, Z.; Wu, Y.; Wang, Y.; Wang, D.; Zhang, W.; Yuan, H.; Ye, J.; Song, X.; Yang, J.; et al. The Hippo effector TAZ promotes cancer stemness by transcriptional activation of SOX2 in head neck squamous cell carcinoma. Cell Death Dis. 2019, 10, 603.
  19. Li, Z.; Wang, Y.; Zhu, Y.; Yuan, C.; Wang, D.; Zhang, W.; Qi, B.; Qiu, J.; Song, X.; Ye, J.; et al. The Hippo transducer TAZ promotes epithelial to mesenchymal transition and cancer stem cell maintenance in oral cancer. Mol. Oncol. 2015, 9, 1091–1105.
  20. Truong, A.B.; Kretz, M.; Ridky, T.W.; Kimmel, R.; Khavari, P.A. p63 regulates proliferation and differentiation of developmentally mature keratinocytes. Genes Dev. 2006, 20, 3185–3197.
  21. Euskirchen, G.; Auerbach, R.K.; Snyder, M. SWI/SNF chromatin-remodeling factors: Multiscale analyses and diverse functions. J. Biol. Chem. 2012, 287, 30897–30905.
  22. Saladi, S.V.; Ross, K.; Karaayvaz, M.; Tata, P.R.; Mou, H.; Rajagopal, J.; Ramaswamy, S.; Ellisen, L.W. ACTL6A Is Co-Amplified with p63 in Squamous Cell Carcinoma to Drive YAP Activation, Regenerative Proliferation, and Poor Prognosis. Cancer Cell 2017, 31, 35–49.
  23. Moleirinho, S.; Chang, N.; Sims, A.H.; Tilston-Lünel, A.M.; Angus, L.; Steele, A.; Boswell, V.; Barnett, S.C.; Ormandy, C.; Faratian, D.; et al. KIBRA exhibits MST-independent functional regulation of the Hippo signaling pathway in mammals. Oncogene 2013, 32, 1821–1830.
  24. Yu, J.; Zheng, Y.; Dong, J.; Klusza, S.; Deng, W.-M.; Pan, D. Kibra functions as a tumor suppressor protein that regulates Hippo signaling in conjunction with Merlin and Expanded. Dev. Cell 2010, 18, 288–299.
  25. Zanconato, F.; Cordenonsi, M.; Piccolo, S. YAP/TAZ at the Roots of Cancer. Cancer Cell 2016, 29, 783–803.
  26. Uhlén, M.; Fagerberg, L.; Hallström, B.M.; Lindskog, C.; Oksvold, P.; Mardinoglu, A.; Sivertsson, Å.; Kampf, C.; Sjöstedt, E.; Asplund, A.; et al. Proteomics. Tissue-based map of the human proteome. Science 2015, 347, 1260419.
  27. Hiemer, S.E.; Zhang, L.; Kartha, V.K.; Packer, T.S.; Almershed, M.; Noonan, V.; Kukuruzinska, M.; Bais, M.V.; Monti, S.; Varelas, X. A YAP/TAZ-Regulated Molecular Signature Is Associated with Oral Squamous Cell Carcinoma. Mol. Cancer Res. MCR 2015, 13, 957–968.
  28. Wang, Y.; Gersten, A.; Moleirinho, S.; Gunn-Moore, F.J.; Reynolds, P.A.; Prystowsky, M.B. Fibroblasts in Head and Neck Squamous Cell Carcinoma Associated With Perineural Invasion Have High-Level Nuclear Yes-Associated Protein (YAP) Expression. Acad. Pathol. 2015, 2, 2374289515616972.
  29. Ge, L.; Smail, M.; Meng, W.; Shyr, Y.; Ye, F.; Fan, K.-H.; Li, X.; Zhou, H.-M.; Bhowmick, N.A. Yes-associated protein expression in head and neck squamous cell carcinoma nodal metastasis. PLoS ONE 2011, 6, e27529.
  30. Wei, Z.; Wang, Y.; Li, Z.; Yuan, C.; Zhang, W.; Wang, D.; Ye, J.; Jiang, H.; Wu, Y.; Cheng, J. Overexpression of Hippo pathway effector TAZ in tongue squamous cell carcinoma: Correlation with clinicopathological features and patients’ prognosis. J. Oral Pathol. Med. Off. Publ. Int. Assoc. Oral Pathol. Am. Acad. Oral Pathol. 2013, 42, 747–754.
  31. Eun, Y.-G.; Lee, D.; Lee, Y.C.; Sohn, B.H.; Kim, E.H.; Yim, S.Y.; Kwon, K.H.; Lee, J.-S. Clinical significance of YAP1 activation in head and neck squamous cell carcinoma. Oncotarget 2017, 8, 111130–111143.
  32. García-Escudero, R.; Segrelles, C.; Dueñas, M.; Pombo, M.; Ballestín, C.; Alonso-Riaño, M.; Nenclares, P.; Álvarez-Rodríguez, R.; Sánchez-Aniceto, G.; Ruíz-Alonso, A.; et al. Overexpression of PIK3CA in head and neck squamous cell carcinoma is associated with poor outcome and activation of the YAP pathway. Oral Oncol. 2018, 79, 55–63.
  33. Pfister, D.G.; Spencer, S.; Adelstein, D.; Adkins, D.; Anzai, Y.; Brizel, D.M.; Bruce, J.Y.; Busse, P.M.; Caudell, J.J.; Cmelak, A.J.; et al. Head and Neck Cancers, Version 2.2020, NCCN Clinical Practice Guidelines in Oncology. J. Natl. Compr. Cancer Netw. 2020, 18, 873–898.
  34. Yoshikawa, K.; Noguchi, K.; Nakano, Y.; Yamamura, M.; Takaoka, K.; Hashimoto-Tamaoki, T.; Kishimoto, H. The Hippo pathway transcriptional co-activator, YAP, confers resistance to cisplatin in human oral squamous cell carcinoma. Int. J. Oncol. 2015, 46, 2364–2370.
  35. Shriwas, O.; Arya, R.; Mohanty, S.; Mohapatra, P.; Kumar, S.; Rath, R.; Kaushik, S.R.; Pahwa, F.; Murmu, K.C.; Majumdar, S.K.D.; et al. RRBP1 rewires cisplatin resistance in oral squamous cell carcinoma by regulating Hippo pathway. Br. J. Cancer 2021, 124, 2004–2016.
  36. Ehsanian, R.; Brown, M.; Lu, H.; Yang, X.P.; Pattatheyil, A.; Yan, B.; Duggal, P.; Chuang, R.; Doondeea, J.; Feller, S.; et al. YAP dysregulation by phosphorylation or ΔNp63-mediated gene repression promotes proliferation, survival and migration in head and neck cancer subsets. Oncogene 2010, 29, 6160–6171.
  37. Albanell, J.; Codony-Servat, J.; Rojo, F.; Del Campo, J.M.; Sauleda, S.; Anido, J.; Raspall, G.; Giralt, J.; Roselló, J.; Nicholson, R.I.; et al. Activated extracellular signal-regulated kinases: Association with epidermal growth factor receptor/transforming growth factor alpha expression in head and neck squamous carcinoma and inhibition by anti-epidermal growth factor receptor treatments. Cancer Res. 2001, 61, 6500–6510.
  38. Søland, T.M.; Husvik, C.; Koppang, H.S.; Boysen, M.; Sandvik, L.; Clausen, O.P.F.; Christoffersen, T.; Bryne, M. A study of phosphorylated ERK1/2 and COX-2 in early stage (T1-T2) oral squamous cell carcinomas. J. Oral Pathol. Med. Off. Publ. Int. Assoc. Oral Pathol. Am. Acad. Oral Pathol. 2008, 37, 535–542.
  39. Judd, N.P.; Winkler, A.E.; Murillo-Sauca, O.; Brotman, J.J.; Law, J.H.; Lewis, J.S.; Dunn, G.P.; Bui, J.D.; Sunwoo, J.B.; Uppaluri, R. ERK1/2 regulation of CD44 modulates oral cancer aggressiveness. Cancer Res. 2012, 72, 365–374.
  40. Uppaluri, R.; Winkler, A.E.; Lin, T.; Law, J.H.; Haughey, B.H.; Nussenbaum, B.; Paniello, R.C.; Rich, J.T.; Diaz, J.A.; Michel, L.P.; et al. Biomarker and Tumor Responses of Oral Cavity Squamous Cell Carcinoma to Trametinib: A Phase II Neoadjuvant Window-of-Opportunity Clinical Trial. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2017, 23, 2186–2194.
  41. Zhao, Y.; Adjei, A.A. The clinical development of MEK inhibitors. Nat. Rev. Clin. Oncol. 2014, 11, 385–400.
  42. Mudianto, T.; Campbell, K.M.; Webb, J.; Zolkind, P.; Skidmore, Z.L.; Riley, R.; Barnell, E.K.; Ozgenc, I.; Giri, T.; Dunn, G.P.; et al. Yap1 Mediates Trametinib Resistance in Head and Neck Squamous Cell Carcinomas. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2021, 27, 2326–2339.
  43. Calses, P.C.; Crawford, J.J.; Lill, J.R.; Dey, A. Hippo Pathway in Cancer: Aberrant Regulation and Therapeutic Opportunities. Trends Cancer 2019, 5, 297–307.
  44. Zanconato, F.; Battilana, G.; Cordenonsi, M.; Piccolo, S. YAP/TAZ as therapeutic targets in cancer. Curr. Opin. Pharmacol. 2016, 29, 26–33.
  45. Bushweller, J.H. Targeting Transcription Factors in Cancer - from Undruggable to Reality. Nat Rev Cancer 2019, 19, 611–624.
  46. Liu-Chittenden, Y.; Huang, B.; Shim, J.S.; Chen, Q.; Lee, S.-J.; Anders, R.A.; Liu, J.O.; Pan, D. Genetic and Pharmacological Disruption of the TEAD-YAP Complex Suppresses the Oncogenic Activity of YAP. Genes Dev 2012, 26, 1300–1305.
  47. Schmidt-Erfurth, U.; Hasan, T. Mechanisms of Action of Photodynamic Therapy with Verteporfin for the Treatment of Age-Related Macular Degeneration. Surv Ophthalmol 2000, 45, 195–214.
  48. Gibault, F.; Corvaisier, M.; Bailly, F.; Huet, G.; Melnyk, P.; Cotelle, P. Non-Photoinduced Biological Properties of Verteporfin. Curr Med Chem 2016, 23, 1171–1184.
  49. Burtness, B.; Harrington, K.J.; Greil, R.; Soulières, D.; Tahara, M.; de Castro, G.; Psyrri, A.; Basté, N.; Neupane, P.; Bratland, Å.; et al. Pembrolizumab alone or with chemotherapy versus cetuximab with chemotherapy for recurrent or metastatic squamous cell carcinoma of the head and neck (KEYNOTE-048): A randomised, open-label, phase 3 study. Lancet Lond. Engl. 2019, 394, 1915–1928.
  50. Tang, T.T.; Konradi, A.W.; Feng, Y.; Peng, X.; Ma, M.; Li, J.; Yu, F.-X.; Guan, K.-L.; Post, L. Small Molecule Inhibitors of TEAD Auto-palmitoylation Selectively Inhibit Proliferation and Tumor Growth of NF2-deficient Mesothelioma. Mol. Cancer Ther. 2021, 20, 986–998.
  51. Wang, W.; Li, N.; Li, X.; Tran, M.K.; Han, X.; Chen, J. Tankyrase Inhibitors Target YAP by Stabilizing Angiomotin Family Proteins. Cell Rep. 2015, 13, 524–532.
  52. Araujo, J.; Logothetis, C. Dasatinib: A potent SRC inhibitor in clinical development for the treatment of solid tumors. Cancer Treat. Rev. 2010, 36, 492–500.
  53. Dunn, L.A.; Riaz, N.; Fury, M.G.; McBride, S.M.; Michel, L.; Lee, N.Y.; Sherman, E.J.; Baxi, S.S.; Haque, S.S.; Katabi, N.; et al. A Phase 1b Study of Cetuximab and BYL719 (Alpelisib) Concurrent with Intensity Modulated Radiation Therapy in Stage III-IVB Head and Neck Squamous Cell Carcinoma. Int. J. Radiat. Oncol. Biol. Phys. 2020, 106, 564–570.
  54. Zhang, B.; Zhang, Y.; Zhang, J.; Liu, P.; Jiao, B.; Wang, Z.; Ren, R. Focal Adhesion Kinase (FAK) Inhibition Synergizes with KRAS G12C Inhibitors in Treating Cancer through the Regulation of the FAK-YAP Signaling. Adv. Sci. Weinh. Baden-Wurtt. Ger. 2021, 8, e2100250.
  55. Pifer, P.; Kumar, M.; Yang, L.; Xie, T.; Frederick, M.; Hefner, A.; Beadle, B.M.; Dhawan, A.; Molkentine, D.; Molkentine, J.; et al. Focal Adhesion Kinase Drives Resistance to Therapy in HPV-Negative Head and Neck Squamous Cell Carcinoma in a p53-Dependent Manner. Int. J. Radiat. Oncol. 2022, 112, e12.
  56. Chan, P.; Han, X.; Zheng, B.; DeRan, M.; Yu, J.; Jarugumilli, G.K.; Deng, H.; Pan, D.; Luo, X.; Wu, X. Autopalmitoylation of TEAD proteins regulates transcriptional output of the Hippo pathway. Nat. Chem. Biol. 2016, 12, 282–289.
  57. Ramms, D.J.; Raimondi, F.; Arang, N.; Herberg, F.W.; Taylor, S.S.; Gutkind, J.S. Gαs-Protein Kinase A (PKA) Pathway Signalopathies: The Emerging Genetic Landscape and Therapeutic Potential of Human Diseases Driven by Aberrant Gαs-PKA Signaling. Pharmacol. Rev. 2021, 73, 155–197.
  58. Santos, R.; Ursu, O.; Gaulton, A.; Bento, A.P.; Donadi, R.S.; Bologa, C.G.; Karlsson, A.; Al-Lazikani, B.; Hersey, A.; Oprea, T.I.; et al. A comprehensive map of molecular drug targets. Nat. Rev. Drug Discov. 2017, 16, 19–34.
  59. Campbell, A.P.; Smrcka, A.V. Targeting G protein-coupled receptor signalling by blocking G proteins. Nat. Rev. Drug Discov. 2018, 17, 789–803.
  60. Wu, V.; Yeerna, H.; Nohata, N.; Chiou, J.; Harismendy, O.; Raimondi, F.; Inoue, A.; Russell, R.B.; Tamayo, P.; Gutkind, J.S. Illuminating the Onco-GPCRome: Novel G protein-coupled receptor-driven oncocrine networks and targets for cancer immunotherapy. J. Biol. Chem. 2019, 294, 11062–11086.
  61. Coles, G.L.; Cristea, S.; Webber, J.T.; Levin, R.S.; Moss, S.M.; He, A.; Sangodkar, J.; Hwang, Y.C.; Arand, J.; Drainas, A.P.; et al. Unbiased Proteomic Profiling Uncovers a Targetable GNAS/PKA/PP2A Axis in Small Cell Lung Cancer Stem Cells. Cancer Cell 2020, 38, 129–143.e7.
  62. Kauko, O.; O’Connor, C.M.; Kulesskiy, E.; Sangodkar, J.; Aakula, A.; Izadmehr, S.; Yetukuri, L.; Yadav, B.; Padzik, A.; Laajala, T.D.; et al. PP2A inhibition is a druggable MEK inhibitor resistance mechanism in KRAS-mutant lung cancer cells. Sci. Transl. Med. 2018, 10, eaaq1093.
  63. Ho, W.S.; Wang, H.; Maggio, D.; Kovach, J.S.; Zhang, Q.; Song, Q.; Marincola, F.M.; Heiss, J.D.; Gilbert, M.R.; Lu, R.; et al. Pharmacologic inhibition of protein phosphatase-2A achieves durable immune-mediated antitumor activity when combined with PD-1 blockade. Nat. Commun. 2018, 9, 2126.
  64. Jung, Y.-S.; Park, J.-I. Wnt signaling in cancer: Therapeutic targeting of Wnt signaling beyond β-catenin and the destruction complex. Exp. Mol. Med. 2020, 52, 183–191.
  65. Morrone, S.; Cheng, Z.; Moon, R.T.; Cong, F.; Xu, W. Crystal structure of a Tankyrase-Axin complex and its implications for Axin turnover and Tankyrase substrate recruitment. Proc. Natl. Acad. Sci. USA 2012, 109, 1500–1505.
  66. Huang, S.-M.A.; Mishina, Y.M.; Liu, S.; Cheung, A.; Stegmeier, F.; Michaud, G.A.; Charlat, O.; Wiellette, E.; Zhang, Y.; Wiessner, S.; et al. Tankyrase inhibition stabilizes axin and antagonizes Wnt signalling. Nature 2009, 461, 614–620.
  67. Azzolin, L.; Zanconato, F.; Bresolin, S.; Forcato, M.; Basso, G.; Bicciato, S.; Cordenonsi, M.; Piccolo, S. Role of TAZ as mediator of Wnt signaling. Cell 2012, 151, 1443–1456.
  68. Juan, W.C.; Hong, W. Targeting the Hippo Signaling Pathway for Tissue Regeneration and Cancer Therapy. Genes 2016, 7, 55.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 1.3K
Revisions: 3 times (View History)
Update Date: 11 May 2022
1000/1000