Next Article in Journal
Cognitive Processes during Recovery: Moving toward Personalized Spine Surgery Outcomes
Next Article in Special Issue
Machine Learning in Predicting Tooth Loss: A Systematic Review and Risk of Bias Assessment
Previous Article in Journal
Stomatognathic System Changes in Obese Patients Undergoing Bariatric Surgery: A Systematic Review
Previous Article in Special Issue
Application of Reactive Oxygen Species in Dental Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Integrating Genetic Alterations and the Hippo Pathway in Head and Neck Squamous Cell Carcinoma for Future Precision Medicine

1
Center of Oral Clinical Examination, Hiroshima University Hospital, Hiroshima 734-8551, Japan
2
Department of Oral Oncology, Graduate School of Biomedical and Health Sciences, Hiroshima University, Hiroshima 734-8553, Japan
3
Department of Oral and Maxillofacial Pathobiology, Graduate School of Biomedical and Health Sciences, Hiroshima University, Hiroshima 734-8553, Japan
4
Moores Cancer Center, University of California, San Diego, CA 92093, USA
5
Department of Pharmacology, University of California, San Diego, CA 92093, USA
*
Author to whom correspondence should be addressed.
J. Pers. Med. 2022, 12(10), 1544; https://doi.org/10.3390/jpm12101544
Submission received: 31 August 2022 / Revised: 14 September 2022 / Accepted: 15 September 2022 / Published: 20 September 2022
(This article belongs to the Special Issue Precision Medicine for Oral Diseases)

Abstract

:
Genetic alterations and dysregulation of signaling pathways are indispensable for the initiation and progression of cancer. Understanding the genetic, molecular, and signaling diversities in cancer patients has driven a dynamic change in cancer therapy. Patients can select a suitable molecularly targeted therapy or immune checkpoint inhibitor based on the driver gene alterations determined by sequencing of cancer tissue. This “precision medicine” approach requires detailed elucidation of the mechanisms connecting genetic alterations of driver genes and aberrant downstream signaling pathways. The regulatory mechanisms of the Hippo pathway and Yes-associated protein/transcriptional co-activator with PDZ binding motif (YAP/TAZ) that have central roles in cancer cell proliferation are not fully understood, reflecting their recent discovery. Nevertheless, emerging evidence has shown that various genetic alterations dysregulate the Hippo pathway and hyperactivate YAP/TAZ in cancers, including head and neck squamous cell carcinoma (HNSCC). Here, we summarize the latest evidence linking genetic alterations and the Hippo pathway in HNSCC, with the aim of contributing to the continued development of precision medicine.

1. Introduction

The Hippo pathway is a tumor-suppressive signaling axis. Its downstream effectors, Yes-associated protein (YAP) and transcriptional co-activator with PDZ binding motif (TAZ), are essential in normal cell growth, tissue growth, and organ size [1]. In mammals, the core Hippo kinase pathway consists of mammalian STE20-like kinase 1 and 2 (MST1/2), large tumor suppressor 1 and 2 (LATS1/2), and their respective adaptor proteins, Salvador homolog 1 (SAV1) and MOB kinase activators 1A and 1B (MOB1A/B) [2,3,4,5,6,7,8,9]. When the Hippo pathway is active, MST1/2 phosphorylates LATS1/2. This, in turn, activates LATS1/2 phosphorylation of YAP/TAZ at five serine residues, resulting in their cytoplasmic retention by binding to 14-3-3 and/or degradation by the ubiquitin-proteasome pathway [10]. In contrast, when the Hippo pathway is inactive, hypophosphorylated YAP/TAZ translocates into the nucleus and interacts with transcription factors, including TEA domain family members (TEAD), to promote the transcription of growth-related genes, including connective growth factor (CTGF) and cysteine-rich angiogenic inducer 61 (CYR61) [11]. Similar to YAP/TAZ, vestigial-like proteins 1-4 (VGLL1-4) also bind to TEAD through their Tondu (TDU) domains. VGLL1-3 contains only one TDU domain at its C-terminus, while VGLL4 has two TDU domains [12,13,14,15]. The function of VGLL is still poorly understood. VGLL1 promotes proliferation of prostate cancer cells [14]. However, VGLL2-4, especially VGLL4, transcriptionally suppress TEAD by competing with YAP [16,17,18,19] (Figure 1).
In addition to the main Hippo components (MST1/2, LATS1/2, SAV1, and MOB1), multiple molecules have been identified as regulators of the Hippo pathway. The angiomotin (AMOT) family, neurofibromin 2 (NF2; also known as Merlin), kidney and brain protein (KIBRA; also known as WWC1), FERM domain-containing 6 (FRMD6), LIM domain-containing protein AJUBA, zonula occludens, and α-catenin are involved in cell-cell junctions and regulate the Hippo pathway [20]. Therefore, the loss of cell-cell adhesion inactivates the Hippo pathway and activates YAP/TAZ [21]. Mitogen-activated protein kinase kinase kinase kinase (MAP4K) family members consist of hematopoietic progenitor kinase 1 (HPK1/MAP4K1), germinal center kinase (GCK/MAP4K2), germinal center kinase-like kinase (GLK/MAP4K3), HPK/GCK-like kinase (HGK/MAP4K4), misshapen-like kinase 1 (MINK1/MAP4K6), and TRAF2 and NCK interacting kinase (TNIK/MAP4K7). These enzymes can phosphorylate the hydrophobic motif (T1079) of LATS1/2, leading to YAP inactivation [22]. TAO kinases (TAOKs) phosphorylate MST1/2 and MAP4Ks, inducing their activation [23,24,25]. Striatin (STRN)-interacting phosphatase and kinase (STRIPAK) is a protein complex composed of PP2A catalytic subunit (PP2AC), scaffolding subunit (PP2AA), and the STRN regulatory subunit. The complex recruits STRN-interacting protein (STRIP1/2), sarcolemmal membrane-associated protein (SLMAP), tumor necrosis factor receptor-associated factor 3-interacting protein 3 (TRAF3IP3), suppressor of IKBKE1 (SIKE1), fibroblast growth factor receptor 1 oncogene partner 2 (FGFR1OP2), cortactin-binding protein 2 (CTTNBP2), CTTNBP2 N-terminal-like protein (CTTNBP2NL), MOB4, GCK family, and cerebral cavernous malformations 3 (CCM3) [26]. Serum or lysophosphatidic acid (LPA) stimulation activates ras homolog family member A (RhoA) and dissociates rhophilin Rho GTPase binding protein 1 (RHPN1) and NF2/KIBRA from STRIPAK, inducing the dephosphorylation of MST1/2 and MAP4K4 by STRIPAK [27]. STRIPAK induces dephosphorylation of MAP4K4, leading to Hippo pathway inactivation and YAP activation [28]. Heat stress induces dephosphorylation and degradation of LATS1/2 by heat shock protein 90 (HSP90) and protein phosphatase 5 (PP5) [29]. Through these aforementioned multiple regulators, the Hippo pathway and YAP/TAZ are controlled by physiological conditions including cell density, mechanical stress, serum through G protein-coupled receptors, and heat stress [29,30] (Figure 1).
The Hippo pathway is dysregulated and YAP/TAZ is predominantly hyperactivated in multiple types of cancer [31]. Genetic alterations are strongly involved in dysregulation of the Hippo pathway. For example, NF2, a suppressor of the Hippo pathway, is frequently mutated or inactivated by gene copy number loss, which leads to the activation of YAP/TAZ in meningiomas, mesotheliomas, and peripheral nerve sheath tumors [31]. GNAQ or GNA11, encoding Gαq and Gα11, respectively, are frequently mutated in uveal melanomas (>80%), which causes YAP/TAZ hyperactivation, resulting in uveal melanoma development [32,33]. Liver kinase B1 (LKB1; also known as STK11) is frequently mutated in non-small-cell lung carcinomas (15–35%) and cervical carcinomas (20%) [34]. LKB1 regulates the microtubule affinity-regulating kinase (MARK) family and scribble homolog (SCRIB) to activate MST1/2 and LATS1/2 [35]. These findings suggest that the somatic mutation of LKB1 drives carcinogenesis through YAP/TAZ activation. In terms of the treatment, however, no drug targeting the Hippo-YAP/TAZ pathway is approved by the Food and Drug Administration (FDA) for the treatment of any types of cancer. Strikingly, despite the oncogenic role of YAP/TAZ in most cancers, the function of the Hippo pathway and YAP/TAZ depends on cancer type. YAP/TAZ plays a tumor suppressive role by inducing integrin alpha 5 and beta 5 in neural/neuroendocrine and RB transcriptional corepressor 1 (RB1)−/− solid cancers, such as retinoblastoma, small cell lung cancer, and neuroendocrine prostate cancer [36]. In estrogen receptor-alpha positive (ERα+) breast cancer, LATS1/2 regulates ERα expression. Loss of LATS1/2 selectively suppresses the proliferation of ERα+, but not ERα, breast cancer cells [37]. Moreover, activated YAP and TEAD enhance VGLL3 transcription. In turn, VGLL3 interacts with TEAD and NCOR2 and inhibits ESR1 transcription from its super enhancer region [38]. LATS1/2 has an oncogenic role by sustaining the Wnt pathway and stem cell function in the intestine, thereby promoting colorectal malignancies [39]. Therefore, cancer type-specific roles of YAP/TAZ should be carefully considered and investigated for precision medicine targeting the Hippo and YAP/TAZ pathways.
Head and neck squamous cell carcinoma (HNSCC) is diagnosed in approximately 54,000 new cases each year in the United States alone, resulting in more than 11,230 deaths [40]. The Cancer Genome Atlas (TCGA) network contains comprehensive information regarding gene alterations in HNSCC. This data have accelerated the understanding of the connections between gene alterations and Hippo pathway dysregulation [41]. Emerging evidence has linked genetic alterations with Hippo pathway dysregulation and aberrant YAP/TAZ activation, which are essential for HNSCC initiation and progression. Here, we introduce the latest discoveries connecting genetic alterations and the Hippo pathway in HNSCC. The aim of this review is to assist the further development of precision medicines for HNSCC patients.

2. Role of the Hippo Pathway and YAP/TAZ in HNSCC

YAP/TAZ protein levels are generally sustained at low levels in the prickled layer but are higher in the basal cells in the normal mucosal epithelium of the head and neck region. YAP/TAZ protein levels increase in precancerous and invasive HNSCC [42]. Nuclear YAP activation is more evident in poorly differentiated HNSCC than in well-differentiated HNSCC [43]. Conditional MOB1A/B double knockout in basal cells initiates squamous cell carcinoma of the tongue within four weeks, suggesting that YAP hyperactivation is sufficient for HNSCC initiation [44]. Several mouse studies have also shown that YAP/TAZ is required for the initiation of basal cell carcinoma and skin squamous cell carcinoma [45,46]. Moreover, loss of heterotrimeric G protein Gαs and protein kinase A (PKA) signaling promotes YAP and glioma-associated oncogene homolog 1 (GLI1) activities, which are sufficient for hair follicle stem cell expansion and basal cell carcinoma initiation [47]. In HNSCC tissue, YAP expression is higher in the invasive front of the tumor than proximal region [48]. Single cell approach revealed partial epithelial-to-mesenchymal transition (EMT) cells localize in the invasive front [49]. Partial-EMT cells also promote cancer stemness [50]. Collectively, these results indicate that YAP/TAZ hyperactivation sustains and expands stem or progenitor cells in the basal cell layer of the mucosal epithelium in the head and neck region, leading to HNSCC initiation and progression.

3. Genetic Alterations and Dysregulation of the Hippo Pathway in HNSCC

YAP1 amplification and minor gene alterations in Hippo components. TCGA includes a comprehensive landscape of somatic genetic alterations in HNSCC, with the second highest incidence of YAP1 gene amplification (4.4% of the cases) and the third highest incidence of WWTR1 (TAZ) gene amplification (8.8%) among all cancers [51]. However, copy number loss or mutations in the core Hippo components (MST1/2, LATS1/2, SAV1, and MOB1A/B) are rare in HNSCC (Figure 2). This suggests that genetic alterations coding for the molecules surrounding the core Hippo pathway components or interacting with YAP/TAZ/TEADs in the nucleus are more important for Hippo pathway dysregulation, leading to YAP/TAZ hyperactivation.
Human papillomavirus (HPV) infection. HPV is an important subset of HNSCC cases globally [52]. Specifically, HPV-16 and HPV-18 are high-risk subtypes associated with many malignancies, including cervical, head and neck, anal, and vulvar cancers [53]. Among the head and neck areas, the oropharynx is the site most frequently affected by HPV, and its incidence is rising [54,55]. E6 and E7 are specific viral oncoproteins that are essential for carcinogenesis through the inhibition of the tumor suppressor proteins p53 and RB [53]. E6/E7 expressing HNSCC is also associated with mammalian target of rapamycin (mTOR) signaling activation [56]. In addition, recent discoveries indicate that YAP/TAZ activation is required for both HPV-negative and HPV-positive HNSCC initiation and progression. The tumor suppressor protein tyrosine phosphatase non-receptor type 14 (PTPN14) is a tumor suppressor inducing YAP inactivation [57,58,59,60]. E7 binds and degrades PTPN14, thereby activating YAP in basal cells in the stratified squamous epithelium, leading to carcinogenesis [61]. E6 homodimerization induces degradation of SCRIB, thereby promoting YAP/TAZ nuclear localization [62]. E6 suppresses YAP degradation and promotes nuclear YAP localization, thereby increasing growth-related genes, including amphiregulin (AREG) and transforming growth factor alpha (TGFA), which encode TGF-α, in cervical cancer [63,64]. Hyperactivation of YAP and E6/E7 synergistically promotes the initiation and progression of cervical cancer in vitro and in vivo [64]. Therefore, YAP/TAZ was hyperactivated in HPV-positive HNSCC cells (Figure 3).
FAT atypical cadherin 1 (FAT1). FAT1 is a member of the FAT family. FAT1 is the second most frequently altered gene (29.8%) after TP53 (70%) [41]. The FAT1 protein is composed of a large extracellular region with cadherin and epidermal growth factor (EGF)-like repeats, a transmembrane region, and an intracellular region [65]. FAT1 can be proteolyzed, and its intracellular domain is translocated into the nucleus [66]. FAT1 has a tumor-suppressive role by interacting with β-catenin and suppressing its nuclear translocation. Thus, frequent somatic mutations in FAT1 cause aberrant Wnt pathway activation in HNSCC [67]. In addition, FAT1 somatic mutations are positively correlated with poor survival of HNSCC patients [68]. Patients harboring truncated FAT1 have a worse prognosis than wild type patients with HPV-negative HNSCC [69]. Bioinformatics and statistical approaches have revealed that FAT1 mutation is prevalent in HPV-negative HNSCC and is associated with human epidermal growth factor receptor 3 (HER3) activation and reduced epidermal growth factor receptor (EGFR) expression, suggesting that FAT1 mutation may confer resistance to EGFR-targeted therapy [70].
Notably, recent emerging evidence has clarified a solid link between FAT1 gene alterations and the Hippo pathway. FAT1 recruits and assembles Hippo components, including NF2, AMOT, MST1/2, SAV1, LATS1/2, and MOB1A/B, and activates them via TAOKs, resulting in YAP inactivation. Thus, frequent deletion or truncation of FAT1 disperses Hippo components and leads to aberrant activation of YAP in HNSCC [43]. Loss of FAT1 also induces a hybrid epithelial to mesenchymal transition (EMT) state with stemness and metastasis via activation of the Ca2+/calmodulin-dependent protein kinase II (CAMK2)-CD44-SRC-YAP axis in mouse and human squamous cell carcinoma [71]. Given the frequent gene alterations of FAT1 in HNSCC, FAT1 gene alterations explain the prevalent YAP activation in HNSCC (Figure 2 and Figure 3).
EGFR. EGFR is an ERBB family tyrosine kinase. EGFR is amplified and highly overexpressed in HNSCC and lung squamous cell carcinoma, frequently mutated and activated in lung adenocarcinoma (LUAC), and mutated, rearranged, and amplified in glioblastoma [41,72,73,74]. Several reports have shown that phosphoinositide-dependent kinase (PDK1) activated by EGFR and phosphoinositide 3-kinase (PI3K) induces Hippo pathway inactivation and YAP activation [75,76]. EGFR, followed by RAS activation, induces SUMOylation of Otubain-2 (OTUB2), thereby promoting deubiquitination of YAP/TAZ and stabilizing them in cancer cells [77]. In addition, recent emerging evidence has revealed that EGFR induces tyrosine phosphorylation of MOB1 at Y95, Y114, and Y117 and inactivates LATS1/2. The resulting activation of YAP/TAZ in HNSCC occurs independent of FAT1 gene alterations and the PI3K-PDK1 axis [78]. In another study, MOB1A/B double knockout in keratin 14 expressing basal cells resulted in the development of tongue squamous cell carcinoma within 4 weeks, suggesting that MOB1A/B knockout is sufficient to suppress LATS1/2 activity and to hyperactivate YAP/TAZ [44]. Given this evidence, tyrosine phosphorylation of MOB1A/B by EGFR may interfere with the physical interactions between LATS1/2 and MST1/2, MAP4Ks, or TAOKs, resulting in LATS1/2 inactivation. Although EGFR has been targeted by cetuximab, an FDA-approved drug for HNSCC patients, monotherapy response rate remains only 10–30% with intrinsic or acquired resistance [79,80,81]. YAP is amplified and overexpressed in cetuximab-resistant HNSCC cell lines [82]. In LUAC harboring frequent EGFR-activating mutations, EGFR-tyrosine kinase inhibitors (EGFR-TKIs) are the main therapeutic drugs. The emergence of YAP reactivation, EGFR-T790M mutations, or activation of other signaling pathways, including MET proto-oncogene, receptor tyrosine kinase (MET), AXL receptor tyrosine kinase (AXL), insulin-like growth factor 1 receptor (IGF1R), interleukin-6 receptor (IL-6R), human epidermal growth factor receptor 2 (HER2), and HER3, are potentially important for resistance to EGFR-TKIs in LUAC [83,84,85]. The observation that LATS1/2 KO by CRISPR/Cas9 confers resistance to EGFR inhibitors in vitro and in vivo [78] indicate that YAP/TAZ reactivation by an unknown mechanism after EGFR inhibitor treatment plays an important role in HNSCC initial resistance or tumor recurrence. In addition, activated YAP/TAZ promotes transcription and expression of amphiregulin (AREG), which in turn acts as a ligand for EGFR [86]. These findings suggest that the sustained EGFR-YAP-AREG axis represents a positive feedback loop in EGFR-altered cancers. Therefore, targeting YAP/TAZ such as by the use of TEAD inhibitors or inhibitors/antibodies for unknown mechanisms in combination with EGFR-targeting therapy may provide novel therapeutic approaches for preventing cancer recurrence and progression (Figure 2, Figure 3 and Figure 4).
Mechanotransduction. Matrix stiffness is an important factor that regulates YAP/TAZ activity. Stiff matrices activate YAP/TAZ through Rho GTPase activity and modification of the actomyosin cytoskeleton [87]. Matrix stiffness activates phospholipase Cγ1 (PLCγ1) to suppress phosphatidylinositol 4,5-bisphosphate (Ptdlns(4,5)P2) and phosphatidic acid, which induce Ras-related GTPase RAP2 activation by PDZ domain-containing guanine nucleotide exchange factor 1 (PDZGEF1) and PDZGEF2. At low stiffness, active RAP2 stimulates MAP4K4, MAP4K6, MAP4K7, and ARHGAP29, leading to LATS1/2 activation and YAP/TAZ inactivation [88].
Tissue inhibitor of metalloproteinase-1 (TIMP-1) is a member of the TIMP family (TIMP-1 to -4) that is overexpressed in many cancers, including HNSCC. TIMP-1 high expression in HNSCC is positively correlated with angiogenesis, EMT, metastasis, and worse prognosis [89]. Previous studies have supported the oncogenic role of TIMP-1 in activating the mitogen-activating protein kinase (MAPK) or PI3K-AKT-mTOR pathway independent of its matrix metalloproteinase (MMP) inhibitory function [90,91,92,93,94]. TIMP-1 forms a complex with CD63 and integrin β1 on collagen, and then activates Src and RhoA, resulting in F-actin assembly that inactivates LATS1/2 and activates YAP/TAZ in many types of cancer cell lines, including HNSCC [95,96]. TIMP-1 released from cancer cells promotes the accumulation of cancer-associated fibroblasts (CAFs) within several cancer types [97]. In contrast, CAFs also require YAP/TAZ activation through the activation of Rho-associated, coiled-coil containing protein kinase (ROCK) and Src, which generates a stiffer matrix in cancer tissue [98]. These results suggest that TIMP-1 induces matrix stiffness with fibrosis and that the stiffness simultaneously induces YAP/TAZ activation in cancer cells via mechanotransduction (Figure 2 and Figure 3).
The VAV family (VAV1-3) are guanosine nucleotide exchange factors (GEFs) that are activated through tyrosine phosphorylation by tyrosine kinases [99]. EGFR phosphorylates and activates VAV guanine nucleotide exchange factor 2 (VAV2), resulting in high levels of GTP-bound (activated) Rac family small GTPase 1 (RAC1) in HNSCC [100]. Moreover, VAV2 is frequently overexpressed in HNSCC, is important for cell proliferation, and sustains an undifferentiated state through activation of Rac1 and RhoA, followed by YAP activation [101]. Interestingly, VAV2 delays EGFR internalization and degradation, thus enhancing downstream signaling pathways [102]. Given that the EGFR-MOB1 axis leads to YAP activation, VAV2 overexpression may be involved in YAP activation in HNSCC by sustaining the EGFR-MOB1 and VAV2-RAC1-YAP axes (Figure 3).
YAP-interacting molecules in the nucleus. Emerging evidence suggests that YAP/TAZ/TEAD interacts with multiple factors in the nucleus. YAP/TEAD2 cooperates with transcription factor E2F and promotes cell cycle gene expression during relapse of KRAS-independent pancreatic ductal adenocarcinoma [103]. Activator protein-1 (AP-1; a dimer of JUN and FOS proteins) and YAP/TAZ/TEAD co-occupy the enhancer region of growth-related genes, resulting in synergistic cell proliferation [104]. YAP/TAZ binds to the promoter region of FOS and enhances its transcription, suggesting a positive feedback loop for AP-1 and YAP/TAZ [105]. The general co-activator bromodomain containing protein 4 (BRD4) physically interacts with YAP/TAZ to recruit YAP/TAZ-bound enhancers close to RNA polymerase II (POL II) at YAP/TAZ-regulated promoters, leading to enhanced transcription of growth-related genes [106]. YAP occupies enhancers or super-enhancers and recruits the mediator complex and the CDK9 elongating kinase, thereby releasing RNA POL II promoter pausing to boost transcriptional elongation [107]. The AT-rich interaction domain 1A (ARID1A)-containing SWI/SNF complex physically interacts with YAP/TAZ and suppresses YAP/TAZ-regulated genes [108]. Cyclin dependent kinase 7 (CDK7) phosphorylates YAP/TAZ at S128 and S90 to avoid ubiquitination for degradation by the CRL4DCAF12 E3 ubiquitin ligase complex [109]. Lysine-specific demethylase 1 (LSD1), a histone demethylase encoded by KDM1A gene, forms the nucleosome remodeling and deacetylase complex with YAP, enhancing the transcription of YAP-target genes [110]. Moreover, YAP/TAZ promotes polyamine levels and cell proliferation by activating transcription of ornithine decarboxylase 1 (Odc1), inducing the hypusination of eukaryotic translation factor 5A (eIF5A) to enhance translation of LSD1 [111]. Co-amplification and overexpression of the p53 family members p63 and ACTL6A, encoding an SWI/SNF subunit, suppresses transcription of WWC1 encoding KIBRA, in turn activating YAP [112] (Figure 3). Collectively, the understanding of the YAP/TAZ regulatory mechanism in the nucleus is expanding, which is also necessary for the advancement of precision medicine targeting YAP/TAZ.
YAP-target genes involved in immune evasion. The interaction between programmed cell death 1 (PD-1) in T cells and programmed cell death 1 ligand 1 (PD-L1) in tumor cells leads to T cell exhaustion. This immune checkpoint mechanism generates a new therapeutic approach that targets PD-1 or PD-L1. Emerging evidence suggests that YAP/TAZ are key regulators of PD-L1 expression in tumor cells. The transcriptional complex of activated YAP and TEAD binds to the PD-L1 promoter in EGFR-TKI-resistant lung adenocarcinoma cell lines [113]. YAP/TAZ binds to TEAD at the promoter region of PD-L1 in human breast cancer cell lines, which is not conserved in mouse cancer cell lines [114]. In contrast, YAP/TEAD binds to the enhancer region of PD-L1 in human non-small cell lung cancer (NSCLC) [115], human malignant pleural mesothelioma [116], and BRAF inhibitor-resistant melanoma cells [117]. In addition, BRD4 enrichment in both promoter and enhancer regions is important for PD-L1 transcription by YAP/TAZ/TEAD in HNSCC [118]. However, another study demonstrated that the loss of LATS1/2 triggers anti-tumor immune responses by releasing nucleic-acid-rich extracellular vesicles [119]. Although PD-L1 regulation by YAP/TAZ has been supported by many studies, further research is required (Figure 3).

4. Therapeutic Approach Targeting the Hippo Pathway and YAP/TAZ

Multiple drugs targeting the Hippo pathway and YAP/TAZ have been identified and examined. Here, we discuss representative therapeutic approaches and representative drugs that are undergoing clinical trials (Figure 5 and Table 1).
Inhibitors targeting YAP or TEAD. Although a few studies have demonstrated drugs that effectively target YAP, antisense oligonucleotides ION-537 reduce YAP protein levels, thereby suppressing tumor growth in a mouse model of hepatocellular cancer and a xenograft model of HNSCC cell lines harboring FAT1 mutation, which is in Phase I clinical trial [120] (Table 1).
Several approaches to target TEAD have been described. The first approach focuses on the nucleocytoplasmic shuttling of TEAD. Osmotic stress induces p38 activation, which binds to TEAD and induces its cytoplasmic translocation. Osmotic stress or p38 overexpression suppresses tumor growth in YAP-driven mesothelioma or uveal melanoma cells [121]. Thus, drugs that induce TEAD cytoplasmic localization can be a therapeutic option for patients with YAP-driven cancer (Figure 3 and Figure 5).
The second approach is to focus on vestigial-like (VGLL) proteins competing with YAP/TAZ for TEAD. Super-TDU, a peptide that mimics VGLL4, potently suppresses gastric tumor growth [122]. Moreover, a TEAD inhibitor in a mouse model, in which the Super-TDU peptide was modified to suppress both YAP-TEAD and TAZ-TEAD interactions, reportedly suppressed E2F transcription and cell proliferation and promoted differentiation through Kruppel-like factor 4 (KLF4) activation in keratinocytes [123]. Verteporfin, a photosensitizer used for the treatment of neovascular macular degeneration, was identified as the first drug that inhibits the YAP-TEAD interaction [124] (Figure 3).
The third approach is to focus on the YAP-binding site of TEAD. Several studies have identified a central pocket in the YAP-binding domain (YBD) of TEAD as a targetable site for small molecule inhibitors, including flufenamic acid, a non-steroidal anti-inflammatory drug (NSAID) [125,126]. Furthermore, TEADs have an intrinsic palmitoylating enzyme-like function, and the auto-palmitoylate cysteine of TEADs is essential for binding to YAP/TAZ [127]. During palmitoylation of TEAD, palmitate is attached via thioester linkage to a cysteine residue, which is supplied exogenously by diet or de novo biosynthesized by fatty acid synthase (FASN) [128]. DC-TEADin02, a potent and selective TEAD auto-methylation inhibitor, was identified in a biochemical study [129]. A small molecule library screen using a TEAD-dependent luciferase reporter identified MGH-CP1 as a TEAD auto-methylation inhibitor [130]. A small molecule that antagonistically binds to the TEAD YBD lipid pocket was shown to potently inhibit TEAD function [131]. VT103 and VT104 are compounds that prevent TEAD auto-methylation and inhibit tumor growth of mesothelioma [132]. MYF-01-37 binds covalently to TEAD and disrupts the YAP-TEAD interaction, enhances EGFR inhibitor-mediated apoptosis, and prevents dormancy in EGFR-mutant non-small cell lung cancer [133]. Currently, IK-930 is undergoing Phase I clinical trial [134] (Table 1). Therefore, targeting YAP itself or the YAP-TEAD interaction is a promising therapy for HNSCC patients (Figure 5).
ROCK inhibitor. TIMP-1, Src, and p53 DNA contact mutations and mechanotransduction hyperactivate RhoA/ROCK1/actomyosin signaling. This, in turn, promotes YAP/TAZ-driven carcinogenesis. ROCK inhibitor selectively antagonizes YAP/TAZ-promoted proliferation in cancer [135] (Figure 5). No clinical trials (solid tumors, unspecified or head and neck) are undergoing.
Inhibitors targeting YAP-interacting molecules. CDK7 selective inhibitor THZ1 reduces CDK7-mediated phosphorylation of YAP/TAZ, thereby inducing its degradation by the CRL4DCAF12 E3 ubiquitin ligase complex [109] (Figure 5). Several drugs (BTX-A51, fadraciclib, samuraciclib, and SY-5609) are in Phase I or II clinical trials (Table 1). The bromodomain and extra-terminal motif (BET) inhibitor JQ1 suppresses YAP/TAZ-driven gene expression and tumorigenesis [106] (Figure 5). Five drugs including BI-894999, BPI-23314, NUV-868, PLX-2853, and SYHA-1801 are in Phase I or II clinical trials (Table 1). LSD1 knockout reportedly suppressed 4NQO-driven mouse tongue oral squamous cell carcinoma and LSD1 inhibition by SP2509 inhibited 4NQO-driven oral squamous cell carcinoma in combination with verteporfin [110]. Inhibition of LSD1 by SP-2577 inhibited liver carcinogenesis in YAP transgenic mice and breast cancer cell growth in an in vivo xenograft model [111] (Figure 5). Three drugs including JBI-802, seclidemstat, and CC 90011 are in Phase II clinical trial (Table 1).
MEK inhibitor. AP-1 and YAP/TAZ synergistically promotes growth-related genes [104], and, given that upstream MAPK-extracellular signal-regulated kinase pathway is frequently activated in HNSCC by EGFR or other receptor tyrosine kinases (RTKs), MEK inhibitors including trametinib can be a potential drug targeting AP-1/YAP/TAZ-driven HNSCC progression. In addition, trametinib, an MAPK kinase (MEK) inhibitor, confers YAP overexpression and growth-related gene expression in HNSCC cell lines and patient-derived xenograft models, suggesting an effective combination therapy targeting YAP and MEK for HNSCC patients [136] (Figure 5). Four drugs including trametinib, binimetinib, cobimetinib, and mirdametinib are in Phase II clinical trials (Table 1).
Src inhibitor. Tyrosine phosphorylation of YAP is also important for its activity, as well as serine/threonine phosphorylation. Tyrosine 357 of YAP is phosphorylated by Yes, Src, and focal adhesion kinase (FAK) [137,138,139,140]. Moreover, Y341, Y357, and Y394 of YAP phosphorylation by Src enhances transcriptional activity, nuclear accumulation, and interaction with TEAD of YAP in skin squamous cell carcinoma [141]. Dasatinib, an Src family kinase inhibitor, suppressed the initiation of YAP1-induced tongue squamous cell carcinoma in a mouse model [44]. Therefore, Src inhibitors are promising drugs that target YAP (Figure 5). Five drugs including dasatinib, repotrectinib, VAL-201, TPX-0046, and elzovantinib are in Phase II or III clinical trials (Table 1).
Metformin. Energy stress is an important regulatory factor for YAP/TAZ. AMP-activated catalytic subunit alpha 2 (AMPKa2; also known as AMPK) activated by energy stress directly phosphorylates YAP Ser 94. This induces the dissociation of YAP-TEAD binding and indirectly inactivates YAP through LATS1/2 inactivation [142]. Metformin-induced energy stress also inactivates LATS1/2 through the inhibition of Rho GTPase and actin cytoskeleton dynamics as well as AMPK activation [143]. AMPK activated by energy stress phosphorylates and increases the protein stability of angiomotin like 2 (AMOTL2), which induces LATS1/2 activation and YAP inactivation [144]. Moreover, YAP is O-GlcNAcylated by O-GlcNAc transferase at serine 109 under high-glucose conditions, which prevents YAP inactivation by LATS1/2. Thus, low glucose levels induce YAP inactivation by loss of O-GlcNAcylation independent of AMPK activation [145]. Metformin, a drug used for patients with type 2 diabetes, as well as other AMPK activators, reportedly suppress tumorigenicity and YAP activity via AMPK activation in vitro and in vivo [142]. In addition, AMPK activation also suppresses mTOR signaling, which is important for premalignancy, initiation, and progression of HNSCC proved in past clinical trials [146,147,148,149,150] (Figure 5). Thus, metformin may prove to be a valuable precision drug.

5. Conclusions and Future Directions

Significant progress has been made in understanding the molecular and genetic mechanisms of the Hippo pathway and YAP/TAZ. Abundant evidence integrating genetic alterations in cancer with the Hippo pathway has emerged. The collective evidence will inform the development of drugs targeting the Hippo pathway specifically and, more generally, the future of precision medicine for HNSCC patients. Simultaneously, drug delivery systems, including antisense oligonucleotides, mRNA, liposomes, and nanoparticles, have dramatically advanced. These developments will aid in the realization of therapeutic approaches targeting the Hippo pathway and YAP/TAZ in cancer cells. Given that EGFR-targeted therapy resistance involves YAP/TAZ reactivation and that TEAD inhibitors may represent effective for YAP/TAZ signaling blockade, targeting the Hippo pathway and YAP/TAZ may be expanded toward combination therapy with other existing drugs, including tyrosine kinase inhibitors, antibodies, and immune checkpoint inhibitors. As the functions of the Hippo pathway and YAP/TAZ differ in many cancer types, it is possible cancer heterogeneity may underlie distinct roles for YAP/TAZ in different HNSCC subpopulations. Further studies, including single-cell approaches, will uncover the comprehensive and heterogeneous activity of YAP/TAZ in HNSCC.
Among the multiple drugs targeting the Hippo pathway, YAP/TAZ/TEAD and their regulators, TEAD inhibitors appear to emerge as the most effective and potent suppressors of YAP/TAZ-driven cancer growth, because TEADs bind YAP/TAZ directly and, hence, are the most downstream effectors of the pathway. Future studies will uncover the effect of TEAD inhibitor treatment alone and advanced combination therapies using TEAD inhibitors for HNSCC patients.
Collectively, integrating genetic alterations and the Hippo pathway has opened a new field of cancer research. This includes the development of analyses or drugs that will alter and improve clinical examination and treatment of HNSCC patients. Further studies will reduce a long distance between emerging biological evidence and current clinical trials and assist the development of precision medicine for HNSCC patients.

Author Contributions

Conceptualization, T.A., J.S.G. and M.K.; writing—original draft preparation, T.A., K.O. and T.S.; writing—review and editing, T.A., S.Y., M.M., J.S.G. and M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by JSPS KAKENHI Grant Number JP20K18477.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The landscape of the Figure 2 was created by c-BioPortal for Cancer Genomics (https://www.cbioportal.org accessed on 19 September 2022) by analyzing the data of the HNSCC TCGA, PanCancer Atlas (2015).

Conflicts of Interest

J.S.G. is a member of the advisory board of Oncoceutics, Domain Therapeutics and Vividion. The authors declare no conflict of interest.

References

  1. Yu, F.X.; Zhao, B.; Guan, K.L. Hippo Pathway in Organ Size Control, Tissue Homeostasis, and Cancer. Cell 2015, 163, 811–828. [Google Scholar] [CrossRef] [PubMed]
  2. Harvey, K.F.; Pfleger, C.M.; Hariharan, I.K. The Drosophila Mst ortholog, hippo, restricts growth and cell proliferation and promotes apoptosis. Cell 2003, 114, 457–467. [Google Scholar] [CrossRef]
  3. Udan, R.S.; Kango-Singh, M.; Nolo, R.; Tao, C.; Halder, G. Hippo promotes proliferation arrest and apoptosis in the Salvador/Warts pathway. Nat. Cell Biol. 2003, 5, 914–920. [Google Scholar] [CrossRef] [PubMed]
  4. Pantalacci, S.; Tapon, N.; Léopold, P. The Salvador partner Hippo promotes apoptosis and cell-cycle exit in Drosophila. Nat. Cell Biol. 2003, 5, 921–927. [Google Scholar] [CrossRef] [PubMed]
  5. Wu, S.; Huang, J.; Dong, J.; Pan, D. hippo encodes a Ste-20 family protein kinase that restricts cell proliferation and promotes apoptosis in conjunction with salvador and warts. Cell 2003, 114, 445–456. [Google Scholar] [CrossRef]
  6. Tapon, N.; Harvey, K.F.; Bell, D.W.; Wahrer, D.C.; Schiripo, T.A.; Haber, D.A.; Hariharan, I.K. salvador Promotes both cell cycle exit and apoptosis in Drosophila and is mutated in human cancer cell lines. Cell 2002, 110, 467–478. [Google Scholar] [CrossRef]
  7. Justice, R.W.; Zilian, O.; Woods, D.F.; Noll, M.; Bryant, P.J. The Drosophila tumor suppressor gene warts encodes a homolog of human myotonic dystrophy kinase and is required for the control of cell shape and proliferation. Genes Dev. 1995, 9, 534–546. [Google Scholar] [CrossRef]
  8. Xu, T.; Wang, W.; Zhang, S.; Stewart, R.A.; Yu, W. Identifying tumor suppressors in genetic mosaics: The Drosophila lats gene encodes a putative protein kinase. Development 1995, 121, 1053–1063. [Google Scholar] [CrossRef]
  9. Kango-Singh, M.; Nolo, R.; Tao, C.; Verstreken, P.; Hiesinger, P.R.; Bellen, H.J.; Halder, G. Shar-pei mediates cell proliferation arrest during imaginal disc growth in Drosophila. Development 2002, 129, 5719–5730. [Google Scholar] [CrossRef]
  10. Zhao, B.; Li, L.; Tumaneng, K.; Wang, C.Y.; Guan, K.L. A coordinated phosphorylation by Lats and CK1 regulates YAP stability through SCF(beta-TRCP). Genes Dev. 2010, 24, 72–85. [Google Scholar] [CrossRef] [Green Version]
  11. Moroishi, T.; Hansen, C.G.; Guan, K.L. The emerging roles of YAP and TAZ in cancer. Nat. Rev. Cancer 2015, 15, 73–79. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, H.H.; Mullett, S.J.; Stewart, A.F. Vgl-4, a novel member of the vestigial-like family of transcription cofactors, regulates alpha1-adrenergic activation of gene expression in cardiac myocytes. J. Biol. Chem. 2004, 279, 30800–30806. [Google Scholar] [CrossRef] [PubMed]
  13. Maeda, T.; Chapman, D.L.; Stewart, A.F. Mammalian vestigial-like 2, a cofactor of TEF-1 and MEF2 transcription factors that promotes skeletal muscle differentiation. J. Biol. Chem. 2002, 277, 48889–48898. [Google Scholar] [CrossRef] [PubMed]
  14. Pobbati, A.V.; Chan, S.W.; Lee, I.; Song, H.; Hong, W. Structural and functional similarity between the Vgll1-TEAD and the YAP-TEAD complexes. Structure 2012, 20, 1135–1140. [Google Scholar] [CrossRef]
  15. Vaudin, P.; Delanoue, R.; Davidson, I.; Silber, J.; Zider, A. TONDU (TDU), a novel human protein related to the product of vestigial (vg) gene of Drosophila melanogaster interacts with vertebrate TEF factors and substitutes for Vg function in wing formation. Development 1999, 126, 4807–4816. [Google Scholar] [CrossRef] [PubMed]
  16. Guo, T.; Lu, Y.; Li, P.; Yin, M.X.; Lv, D.; Zhang, W.; Wang, H.; Zhou, Z.; Ji, H.; Zhao, Y.; et al. A novel partner of Scalloped regulates Hippo signaling via antagonizing Scalloped-Yorkie activity. Cell Res. 2013, 23, 1201–1214. [Google Scholar] [CrossRef]
  17. Koontz, L.M.; Liu-Chittenden, Y.; Yin, F.; Zheng, Y.; Yu, J.; Huang, B.; Chen, Q.; Wu, S.; Pan, D. The Hippo effector Yorkie controls normal tissue growth by antagonizing scalloped-mediated default repression. Dev. Cell 2013, 25, 388–401. [Google Scholar] [CrossRef]
  18. Zhang, W.; Gao, Y.; Li, P.; Shi, Z.; Guo, T.; Li, F.; Han, X.; Feng, Y.; Zheng, C.; Wang, Z.; et al. VGLL4 functions as a new tumor suppressor in lung cancer by negatively regulating the YAP-TEAD transcriptional complex. Cell Res. 2014, 24, 331–343. [Google Scholar] [CrossRef]
  19. Zhang, Y.; Shen, H.; Withers, H.G.; Yang, N.; Denson, K.E.; Mussell, A.L.; Truskinovsky, A.; Fan, Q.; Gelman, I.H.; Frangou, C.; et al. VGLL4 Selectively Represses YAP-Dependent Gene Induction and Tumorigenic Phenotypes in Breast Cancer. Sci. Rep. 2017, 7, 6190. [Google Scholar] [CrossRef]
  20. Moya, I.M.; Halder, G. Hippo-YAP/TAZ signalling in organ regeneration and regenerative medicine. Nat. Rev. Mol. Cell Biol. 2019, 20, 211–226. [Google Scholar] [CrossRef]
  21. Zhao, B.; Li, L.; Lu, Q.; Wang, L.H.; Liu, C.Y.; Lei, Q.; Guan, K.L. Angiomotin is a novel Hippo pathway component that inhibits YAP oncoprotein. Genes Dev. 2011, 25, 51–63. [Google Scholar] [CrossRef] [PubMed]
  22. Meng, Z.; Moroishi, T.; Mottier-Pavie, V.; Plouffe, S.W.; Hansen, C.G.; Hong, A.W.; Park, H.W.; Mo, J.S.; Lu, W.; Lu, S.; et al. MAP4K family kinases act in parallel to MST1/2 to activate LATS1/2 in the Hippo pathway. Nat. Commun. 2015, 6, 8357. [Google Scholar] [CrossRef] [PubMed]
  23. Boggiano, J.C.; Vanderzalm, P.J.; Fehon, R.G. Tao-1 phosphorylates Hippo/MST kinases to regulate the Hippo-Salvador-Warts tumor suppressor pathway. Dev. Cell 2011, 21, 888–895. [Google Scholar] [CrossRef] [PubMed]
  24. Poon, C.L.; Lin, J.I.; Zhang, X.; Harvey, K.F. The sterile 20-like kinase Tao-1 controls tissue growth by regulating the Salvador-Warts-Hippo pathway. Dev. Cell 2011, 21, 896–906. [Google Scholar] [CrossRef] [PubMed]
  25. Li, Q.; Nirala, N.K.; Nie, Y.; Chen, H.J.; Ostroff, G.; Mao, J.; Wang, Q.; Xu, L.; Ip, Y.T. Ingestion of Food Particles Regulates the Mechanosensing Misshapen-Yorkie Pathway in Drosophila Intestinal Growth. Dev. Cell 2018, 45, 433–449.e6. [Google Scholar] [CrossRef]
  26. Shi, Z.; Jiao, S.; Zhou, Z. STRIPAK complexes in cell signaling and cancer. Oncogene 2016, 35, 4549–4557. [Google Scholar] [CrossRef]
  27. Chen, R.; Xie, R.; Meng, Z.; Ma, S.; Guan, K.L. STRIPAK integrates upstream signals to initiate the Hippo kinase cascade. Nat. Cell Biol. 2019, 21, 1565–1577. [Google Scholar] [CrossRef]
  28. Kim, J.W.; Berrios, C.; Kim, M.; Schade, A.E.; Adelmant, G.; Yeerna, H.; Damato, E.; Iniguez, A.B.; Florens, L.; Washburn, M.P.; et al. STRIPAK directs PP2A activity toward MAP4K4 to promote oncogenic transformation of human cells. Elife 2020, 9, e53003. [Google Scholar] [CrossRef]
  29. Luo, M.; Meng, Z.; Moroishi, T.; Lin, K.C.; Shen, G.; Mo, F.; Shao, B.; Wei, X.; Zhang, P.; Wei, Y.; et al. Heat stress activates YAP/TAZ to induce the heat shock transcriptome. Nat. Cell Biol. 2020, 22, 1447–1459. [Google Scholar] [CrossRef]
  30. Yu, F.X.; Zhao, B.; Panupinthu, N.; Jewell, J.L.; Lian, I.; Wang, L.H.; Zhao, J.; Yuan, H.; Tumaneng, K.; Li, H.; et al. Regulation of the Hippo-YAP pathway by G-protein-coupled receptor signaling. Cell 2012, 150, 780–791. [Google Scholar] [CrossRef] [Green Version]
  31. Harvey, K.F.; Zhang, X.; Thomas, D.M. The Hippo pathway and human cancer. Nat. Rev. Cancer 2013, 13, 246–257. [Google Scholar] [CrossRef] [PubMed]
  32. Feng, X.; Degese, M.S.; Iglesias-Bartolome, R.; Vaque, J.P.; Molinolo, A.A.; Rodrigues, M.; Zaidi, M.R.; Ksander, B.R.; Merlino, G.; Sodhi, A.; et al. Hippo-independent activation of YAP by the GNAQ uveal melanoma oncogene through a trio-regulated rho GTPase signaling circuitry. Cancer Cell 2014, 25, 831–845. [Google Scholar] [CrossRef] [PubMed]
  33. Yu, F.X.; Luo, J.; Mo, J.S.; Liu, G.; Kim, Y.C.; Meng, Z.; Zhao, L.; Peyman, G.; Ouyang, H.; Jiang, W.; et al. Mutant Gq/11 promote uveal melanoma tumorigenesis by activating YAP. Cancer Cell 2014, 25, 822–830. [Google Scholar] [CrossRef]
  34. Shackelford, D.B.; Shaw, R.J. The LKB1-AMPK pathway: Metabolism and growth control in tumour suppression. Nat. Rev. Cancer 2009, 9, 563–575. [Google Scholar] [CrossRef] [PubMed]
  35. Mohseni, M.; Sun, J.; Lau, A.; Curtis, S.; Goldsmith, J.; Fox, V.L.; Wei, C.; Frazier, M.; Samson, O.; Wong, K.K.; et al. A genetic screen identifies an LKB1-MARK signalling axis controlling the Hippo-YAP pathway. Nat. Cell Biol. 2014, 16, 108–117. [Google Scholar] [CrossRef]
  36. Pearson, J.D.; Huang, K.; Pacal, M.; McCurdy, S.R.; Lu, S.; Aubry, A.; Yu, T.; Wadosky, K.M.; Zhang, L.; Wang, T.; et al. Binary pan-cancer classes with distinct vulnerabilities defined by pro- or anti-cancer YAP/TEAD activity. Cancer Cell 2021, 39, 1115–1134.e12. [Google Scholar] [CrossRef] [PubMed]
  37. Ma, S.; Wu, Z.; Yang, F.; Zhang, J.; Johnson, R.L.; Rosenfeld, M.G.; Guan, K.L. Hippo signalling maintains ER expression and ER+ breast cancer growth. Nature 2021, 591, E1–E10. [Google Scholar] [CrossRef]
  38. Ma, S.; Tang, T.; Probst, G.; Konradi, A.; Jin, C.; Li, F.; Silvio Gutkind, J.; Fu, X.D.; Guan, K.L. Transcriptional repression of estrogen receptor alpha by YAP reveals the Hippo pathway as therapeutic target for ER+ breast cancer. Nat. Commun. 2022, 13, 1061. [Google Scholar] [CrossRef]
  39. Cheung, P.; Xiol, J.; Dill, M.T.; Yuan, W.C.; Panero, R.; Roper, J.; Osorio, F.G.; Maglic, D.; Li, Q.; Gurung, B.; et al. Regenerative Reprogramming of the Intestinal Stem Cell State via Hippo Signaling Suppresses Metastatic Colorectal Cancer. Cell Stem Cell 2020, 27, 590–604.e9. [Google Scholar] [CrossRef]
  40. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  41. Cancer Genome Atlas Research Network. Comprehensive genomic characterization of head and neck squamous cell carcinomas. Nature 2015, 517, 576–582. [Google Scholar] [CrossRef] [PubMed]
  42. Shin, E.; Kim, J. The potential role of YAP in head and neck squamous cell carcinoma. Exp. Mol. Med. 2020, 52, 1264–1274. [Google Scholar] [CrossRef] [PubMed]
  43. 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. [Google Scholar] [CrossRef] [PubMed]
  44. 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. [Google Scholar] [CrossRef] [PubMed]
  45. Debaugnies, M.; Sánchez-Danés, A.; Rorive, S.; Raphaël, M.; Liagre, M.; Parent, M.A.; Brisebarre, A.; Salmon, I.; Blanpain, C. YAP and TAZ are essential for basal and squamous cell carcinoma initiation. EMBO Rep. 2018, 19, e45809. [Google Scholar] [CrossRef]
  46. Schlegelmilch, K.; Mohseni, M.; Kirak, O.; Pruszak, J.; Rodriguez, J.R.; Zhou, D.; Kreger, B.T.; Vasioukhin, V.; Avruch, J.; Brummelkamp, T.R.; et al. Yap1 acts downstream of α-catenin to control epidermal proliferation. Cell 2011, 144, 782–795. [Google Scholar] [CrossRef]
  47. Iglesias-Bartolome, R.; Torres, D.; Marone, R.; Feng, X.; Martin, D.; Simaan, M.; Chen, M.; Weinstein, L.S.; Taylor, S.S.; Molinolo, A.A.; et al. Inactivation of a Gα(s)-PKA tumour suppressor pathway in skin stem cells initiates basal-cell carcinogenesis. Nat. Cell Biol. 2015, 17, 793–803. [Google Scholar] [CrossRef]
  48. 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. [Google Scholar] [CrossRef]
  49. Puram, S.V.; Tirosh, I.; Parikh, A.S.; Patel, A.P.; Yizhak, K.; Gillespie, S.; Rodman, C.; Luo, C.L.; Mroz, E.A.; Emerick, K.S.; et al. Single-Cell Transcriptomic Analysis of Primary and Metastatic Tumor Ecosystems in Head and Neck Cancer. Cell 2017, 171, 1611–1624.e24. [Google Scholar] [CrossRef]
  50. Pal, A.; Barrett, T.F.; Paolini, R.; Parikh, A.; Puram, S.V. Partial EMT in head and neck cancer biology: A spectrum instead of a switch. Oncogene 2021, 40, 5049–5065. [Google Scholar] [CrossRef]
  51. Leemans, C.R.; Snijders, P.J.F.; Brakenhoff, R.H. The molecular landscape of head and neck cancer. Nat. Rev. Cancer 2018, 18, 269–282. [Google Scholar] [CrossRef] [PubMed]
  52. Gillison, M.L.; Chaturvedi, A.K.; Anderson, W.F.; Fakhry, C. Epidemiology of Human Papillomavirus-Positive Head and Neck Squamous Cell Carcinoma. J. Clin. Oncol. 2015, 33, 3235–3242. [Google Scholar] [CrossRef] [PubMed]
  53. zur Hausen, H. Papillomaviruses and cancer: From basic studies to clinical application. Nat. Rev. Cancer 2002, 2, 342–350. [Google Scholar] [CrossRef] [PubMed]
  54. Chaturvedi, A.K.; Engels, E.A.; Pfeiffer, R.M.; Hernandez, B.Y.; Xiao, W.; Kim, E.; Jiang, B.; Goodman, M.T.; Sibug-Saber, M.; Cozen, W.; et al. Human papillomavirus and rising oropharyngeal cancer incidence in the United States. J. Clin. Oncol. 2011, 29, 4294–4301. [Google Scholar] [CrossRef] [PubMed]
  55. Ang, K.K.; Harris, J.; Wheeler, R.; Weber, R.; Rosenthal, D.I.; Nguyen-Tân, P.F.; Westra, W.H.; Chung, C.H.; Jordan, R.C.; Lu, C.; et al. Human papillomavirus and survival of patients with oropharyngeal cancer. N. Engl. J. Med. 2010, 363, 24–35. [Google Scholar] [CrossRef]
  56. Callejas-Valera, J.L.; Iglesias-Bartolome, R.; Amornphimoltham, P.; Palacios-Garcia, J.; Martin, D.; Califano, J.A.; Molinolo, A.A.; Gutkind, J.S. mTOR inhibition prevents rapid-onset of carcinogen-induced malignancies in a novel inducible HPV-16 E6/E7 mouse model. Carcinogenesis 2016, 37, 1014–1025. [Google Scholar] [CrossRef]
  57. Wang, W.; Huang, J.; Wang, X.; Yuan, J.; Li, X.; Feng, L.; Park, J.I.; Chen, J. PTPN14 is required for the density-dependent control of YAP1. Genes Dev. 2012, 26, 1959–1971. [Google Scholar] [CrossRef]
  58. Poernbacher, I.; Baumgartner, R.; Marada, S.K.; Edwards, K.; Stocker, H. Drosophila Pez acts in Hippo signaling to restrict intestinal stem cell proliferation. Curr. Biol. 2012, 22, 389–396. [Google Scholar] [CrossRef]
  59. Mello, S.S.; Valente, L.J.; Raj, N.; Seoane, J.A.; Flowers, B.M.; McClendon, J.; Bieging-Rolett, K.T.; Lee, J.; Ivanochko, D.; Kozak, M.M.; et al. A p53 Super-tumor Suppressor Reveals a Tumor Suppressive p53-Ptpn14-Yap Axis in Pancreatic Cancer. Cancer Cell 2017, 32, 460–473.e6. [Google Scholar] [CrossRef]
  60. Knight, J.F.; Sung, V.Y.C.; Kuzmin, E.; Couzens, A.L.; de Verteuil, D.A.; Ratcliffe, C.D.H.; Coelho, P.P.; Johnson, R.M.; Samavarchi-Tehrani, P.; Gruosso, T.; et al. KIBRA (WWC1) Is a Metastasis Suppressor Gene Affected by Chromosome 5q Loss in Triple-Negative Breast Cancer. Cell Rep. 2018, 22, 3191–3205. [Google Scholar] [CrossRef] [Green Version]
  61. Hatterschide, J.; Castagnino, P.; Kim, H.W.; Sperry, S.M.; Montone, K.T.; Basu, D.; White, E.A. YAP1 activation by human papillomavirus E7 promotes basal cell identity in squamous epithelia. elife 2022, 11, e75466. [Google Scholar] [CrossRef] [PubMed]
  62. Messa, L.; Celegato, M.; Bertagnin, C.; Mercorelli, B.; Alvisi, G.; Banks, L.; Palù, G.; Loregian, A. The Dimeric Form of HPV16 E6 Is Crucial to Drive YAP/TAZ Upregulation through the Targeting of hScrib. Cancers 2021, 13, 4083. [Google Scholar] [CrossRef] [PubMed]
  63. He, C.; Mao, D.; Hua, G.; Lv, X.; Chen, X.; Angeletti, P.C.; Dong, J.; Remmenga, S.W.; Rodabaugh, K.J.; Zhou, J.; et al. The Hippo/YAP pathway interacts with EGFR signaling and HPV oncoproteins to regulate cervical cancer progression. EMBO Mol. Med. 2015, 7, 1426–1449. [Google Scholar] [CrossRef] [PubMed]
  64. Webb Strickland, S.; Brimer, N.; Lyons, C.; Vande Pol, S.B. Human Papillomavirus E6 interaction with cellular PDZ domain proteins modulates YAP nuclear localization. Virology 2018, 516, 127–138. [Google Scholar] [CrossRef]
  65. Dunne, J.; Hanby, A.M.; Poulsom, R.; Jones, T.A.; Sheer, D.; Chin, W.G.; Da, S.M.; Zhao, Q.; Beverley, P.C.; Owen, M.J. Molecular cloning and tissue expression of FAT, the human homologue of the Drosophila fat gene that is located on chromosome 4q34-q35 and encodes a putative adhesion molecule. Genomics 1995, 30, 207–223. [Google Scholar] [CrossRef]
  66. Magg, T.; Schreiner, D.; Solis, G.P.; Bade, E.G.; Hofer, H.W. Processing of the human protocadherin Fat1 and translocation of its cytoplasmic domain to the nucleus. Exp. Cell Res. 2005, 307, 100–108. [Google Scholar] [CrossRef]
  67. Morris, L.G.; 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. [Google Scholar] [CrossRef]
  68. Lin, S.C.; Lin, L.H.; Yu, S.Y.; Kao, S.Y.; Chang, K.W.; Cheng, H.W.; Liu, C.J. FAT1 somatic mutations in head and neck carcinoma are associated with tumor progression and survival. Carcinogenesis 2018, 39, 1320–1330. [Google Scholar] [CrossRef]
  69. Huang, C.; Chen, L.; Savage, S.R.; Eguez, R.V.; Dou, Y.; Li, Y.; da Veiga Leprevost, F.; Jaehnig, E.J.; Lei, J.T.; Wen, B.; et al. Proteogenomic insights into the biology and treatment of HPV-negative head and neck squamous cell carcinoma. Cancer Cell 2021, 39, 361–379.e16. [Google Scholar] [CrossRef]
  70. Chen, Z.; Zhang, C.; Chen, J.; Wang, D.; Tu, J.; Van Waes, C.; Saba, N.F.; Chen, Z.G.; Chen, Z. The Proteomic Landscape of Growth Factor Signaling Networks Associated with FAT1 Mutations in Head and Neck Cancers. Cancer Res. 2021, 81, 4402–4416. [Google Scholar] [CrossRef]
  71. 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. [Google Scholar] [CrossRef] [PubMed]
  72. Cancer Genome Atlas Research Network. Comprehensive genomic characterization of squamous cell lung cancers. Nature 2012, 489, 519–525. [Google Scholar] [CrossRef] [PubMed]
  73. Comprehensive molecular profiling of lung adenocarcinoma. Nature 2014, 511, 543–550. [CrossRef] [PubMed]
  74. Brennan, C.W.; Verhaak, R.G.; McKenna, A.; Campos, B.; Noushmehr, H.; Salama, S.R.; Zheng, S.; Chakravarty, D.; Sanborn, J.Z.; Berman, S.H.; et al. The somatic genomic landscape of glioblastoma. Cell 2013, 155, 462–477. [Google Scholar] [CrossRef]
  75. Xia, H.; Dai, X.; Yu, H.; Zhou, S.; Fan, Z.; Wei, G.; Tang, Q.; Gong, Q.; Bi, F. EGFR-PI3K-PDK1 pathway regulates YAP signaling in hepatocellular carcinoma: The mechanism and its implications in targeted therapy. Cell Death Dis. 2018, 9, 269. [Google Scholar] [CrossRef] [PubMed]
  76. 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. [Google Scholar] [CrossRef]
  77. Zhang, Z.; Du, J.; Wang, S.; Shao, L.; Jin, K.; Li, F.; Wei, B.; Ding, W.; Fu, P.; van Dam, H.; et al. OTUB2 Promotes Cancer Metastasis via Hippo-Independent Activation of YAP and TAZ. Mol. Cell 2019, 73, 7–21.e7. [Google Scholar] [CrossRef]
  78. 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. [Google Scholar] [CrossRef]
  79. Bonner, J.A.; Harari, P.M.; Giralt, J.; Azarnia, N.; Shin, D.M.; Cohen, R.B.; Jones, C.U.; Sur, R.; Raben, D.; Jassem, J.; et al. Radiotherapy plus cetuximab for squamous-cell carcinoma of the head and neck. N. Engl. J. Med. 2006, 354, 567–578. [Google Scholar] [CrossRef]
  80. Vermorken, J.B.; Mesia, R.; Rivera, F.; Remenar, E.; Kawecki, A.; Rottey, S.; Erfan, J.; Zabolotnyy, D.; Kienzer, H.R.; Cupissol, D.; et al. Platinum-based chemotherapy plus cetuximab in head and neck cancer. N. Engl. J. Med. 2008, 359, 1116–1127. [Google Scholar] [CrossRef] [Green Version]
  81. Byeon, H.K.; Ku, M.; Yang, J. Beyond EGFR inhibition: Multilateral combat strategies to stop the progression of head and neck cancer. Exp. Mol. Med. 2019, 51, 1–14. [Google Scholar] [CrossRef] [PubMed]
  82. Jerhammar, F.; Johansson, A.C.; Ceder, R.; Welander, J.; Jansson, A.; Grafström, R.C.; Söderkvist, P.; Roberg, K. YAP1 is a potential biomarker for cetuximab resistance in head and neck cancer. Oral Oncol. 2014, 50, 832–839. [Google Scholar] [CrossRef]
  83. Rotow, J.; Bivona, T.G. Understanding and targeting resistance mechanisms in NSCLC. Nat. Rev. Cancer 2017, 17, 637–658. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, Y.; Dong, Q.; Zhang, Q.; Li, Z.; Wang, E.; Qiu, X. Overexpression of yes-associated protein contributes to progression and poor prognosis of non-small-cell lung cancer. Cancer Sci. 2010, 101, 1279–1285. [Google Scholar] [CrossRef] [PubMed]
  85. Hsu, P.C.; You, B.; Yang, Y.L.; Zhang, W.Q.; Wang, Y.C.; Xu, Z.; Dai, Y.; Liu, S.; Yang, C.T.; Li, H.; et al. YAP promotes erlotinib resistance in human non-small cell lung cancer cells. Oncotarget 2016, 7, 51922–51933. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, J.; Ji, J.Y.; Yu, M.; Overholtzer, M.; Smolen, G.A.; Wang, R.; Brugge, J.S.; Dyson, N.J.; Haber, D.A. YAP-dependent induction of amphiregulin identifies a non-cell-autonomous component of the Hippo pathway. Nat. Cell Biol. 2009, 11, 1444–1450. [Google Scholar] [CrossRef]
  87. Dupont, S.; Morsut, L.; Aragona, M.; Enzo, E.; Giulitti, S.; Cordenonsi, M.; Zanconato, F.; Le Digabel, J.; Forcato, M.; Bicciato, S.; et al. Role of YAP/TAZ in mechanotransduction. Nature 2011, 474, 179–183. [Google Scholar] [CrossRef]
  88. Meng, Z.; Qiu, Y.; Lin, K.C.; Kumar, A.; Placone, J.K.; Fang, C.; Wang, K.C.; Lu, S.; Pan, M.; Hong, A.W.; et al. RAP2 mediates mechanoresponses of the Hippo pathway. Nature 2018, 560, 655–660. [Google Scholar] [CrossRef]
  89. Cai, M.; Zheng, Z.; Bai, Z.; Ouyang, K.; Wu, Q.; Xu, S.; Huang, L.; Jiang, Y.; Wang, L.; Gao, J.; et al. Overexpression of angiogenic factors and matrix metalloproteinases in the saliva of oral squamous cell carcinoma patients: Potential non-invasive diagnostic and therapeutic biomarkers. BMC Cancer 2022, 22, 530. [Google Scholar] [CrossRef]
  90. Xia, Y.; Yeddula, N.; Leblanc, M.; Ke, E.; Zhang, Y.; Oldfield, E.; Shaw, R.J.; Verma, I.M. Reduced cell proliferation by IKK2 depletion in a mouse lung-cancer model. Nat. Cell Biol. 2012, 14, 257–265. [Google Scholar] [CrossRef] [Green Version]
  91. Toricelli, M.; Melo, F.H.; Peres, G.B.; Silva, D.C.; Jasiulionis, M.G. Timp1 interacts with beta-1 integrin and CD63 along melanoma genesis and confers anoikis resistance by activating PI3-K signaling pathway independently of Akt phosphorylation. Mol. Cancer 2013, 12, 1095. [Google Scholar] [CrossRef] [PubMed]
  92. Song, G.; Xu, S.; Zhang, H.; Wang, Y.; Xiao, C.; Jiang, T.; Wu, L.; Zhang, T.; Sun, X.; Zhong, L.; et al. TIMP1 is a prognostic marker for the progression and metastasis of colon cancer through FAK-PI3K/AKT and MAPK pathway. J. Exp. Clin. Cancer Res. 2016, 35, 148. [Google Scholar] [CrossRef] [PubMed]
  93. Jung, K.K.; Liu, X.W.; Chirco, R.; Fridman, R.; Kim, H.R. Identification of CD63 as a tissue inhibitor of metalloproteinase-1 interacting cell surface protein. EMBO J. 2006, 25, 3934–3942. [Google Scholar] [CrossRef] [PubMed]
  94. Chirco, R.; Liu, X.W.; Jung, K.K.; Kim, H.R. Novel functions of TIMPs in cell signaling. Cancer Metastasis Rev. 2006, 25, 99–113. [Google Scholar] [CrossRef]
  95. Ando, T.; Charindra, D.; Shrestha, M.; Umehara, H.; Ogawa, I.; Miyauchi, M.; Takata, T. Tissue inhibitor of metalloproteinase-1 promotes cell proliferation through YAP/TAZ activation in cancer. Oncogene 2018, 37, 263–270. [Google Scholar] [CrossRef]
  96. Shrestha, M.; Ando, T.; Chea, C.; Sakamoto, S.; Nishisaka, T.; Ogawa, I.; Miyauchi, M.; Takata, T. The transition of tissue inhibitor of metalloproteinases from −4 to −1 induces aggressive behavior and poor patient survival in dedifferentiated liposarcoma via YAP/TAZ activation. Carcinogenesis 2019, 40, 1288–1297. [Google Scholar] [CrossRef]
  97. Gong, Y.; Scott, E.; Lu, R.; Xu, Y.; Oh, W.K.; Yu, Q. TIMP-1 promotes accumulation of cancer associated fibroblasts and cancer progression. PLoS ONE 2013, 8, e77366. [Google Scholar] [CrossRef]
  98. Calvo, F.; Ege, N.; Grande-Garcia, A.; Hooper, S.; Jenkins, R.P.; Chaudhry, S.I.; Harrington, K.; Williamson, P.; Moeendarbary, E.; Charras, G.; et al. Mechanotransduction and YAP-dependent matrix remodelling is required for the generation and maintenance of cancer-associated fibroblasts. Nat. Cell Biol. 2013, 15, 637–646. [Google Scholar] [CrossRef]
  99. Rodríguez-Fdez, S.; Bustelo, X.R. The Vav GEF Family: An Evolutionary and Functional Perspective. Cells 2019, 8, 465. [Google Scholar] [CrossRef]
  100. Patel, V.; Rosenfeldt, H.M.; Lyons, R.; Servitja, J.M.; Bustelo, X.R.; Siroff, M.; Gutkind, J.S. Persistent activation of Rac1 in squamous carcinomas of the head and neck: Evidence for an EGFR/Vav2 signaling axis involved in cell invasion. Carcinogenesis 2007, 28, 1145–1152. [Google Scholar] [CrossRef] [Green Version]
  101. Lorenzo-Martín, L.F.; Fernández-Parejo, N.; Menacho-Márquez, M.; Rodríguez-Fdez, S.; Robles-Valero, J.; Zumalave, S.; Fabbiano, S.; Pascual, G.; García-Pedrero, J.M.; Abad, A.; et al. VAV2 signaling promotes regenerative proliferation in both cutaneous and head and neck squamous cell carcinoma. Nat. Commun. 2020, 11, 4788. [Google Scholar] [CrossRef] [PubMed]
  102. Thalappilly, S.; Soubeyran, P.; Iovanna, J.L.; Dusetti, N.J. VAV2 regulates epidermal growth factor receptor endocytosis and degradation. Oncogene 2010, 29, 2528–2539. [Google Scholar] [CrossRef] [PubMed]
  103. Kapoor, A.; Yao, W.; Ying, H.; Hua, S.; Liewen, A.; Wang, Q.; Zhong, Y.; Wu, C.J.; Sadanandam, A.; Hu, B.; et al. Yap1 activation enables bypass of oncogenic Kras addiction in pancreatic cancer. Cell 2014, 158, 185–197. [Google Scholar] [CrossRef]
  104. Zanconato, F.; Forcato, M.; Battilana, G.; Azzolin, L.; Quaranta, E.; Bodega, B.; Rosato, A.; Bicciato, S.; Cordenonsi, M.; Piccolo, S. Genome-wide association between YAP/TAZ/TEAD and AP-1 at enhancers drives oncogenic growth. Nat. Cell Biol. 2015, 17, 1218–1227. [Google Scholar] [CrossRef]
  105. Koo, J.H.; Plouffe, S.W.; Meng, Z.; Lee, D.H.; Yang, D.; Lim, D.S.; Wang, C.Y.; Guan, K.L. Induction of AP-1 by YAP/TAZ contributes to cell proliferation and organ growth. Genes Dev. 2020, 34, 72–86. [Google Scholar] [CrossRef]
  106. Zanconato, F.; Battilana, G.; Forcato, M.; Filippi, L.; Azzolin, L.; Manfrin, A.; Quaranta, E.; Di Biagio, D.; Sigismondo, G.; Guzzardo, V.; et al. Transcriptional addiction in cancer cells is mediated by YAP/TAZ through BRD4. Nat. Med. 2018, 24, 1599–1610. [Google Scholar] [CrossRef] [PubMed]
  107. Galli, G.G.; Carrara, M.; Yuan, W.C.; Valdes-Quezada, C.; Gurung, B.; Pepe-Mooney, B.; Zhang, T.; Geeven, G.; Gray, N.S.; de Laat, W.; et al. YAP Drives Growth by Controlling Transcriptional Pause Release from Dynamic Enhancers. Mol. Cell 2015, 60, 328–337. [Google Scholar] [CrossRef] [PubMed]
  108. Chang, L.; Azzolin, L.; Di Biagio, D.; Zanconato, F.; Battilana, G.; Lucon Xiccato, R.; Aragona, M.; Giulitti, S.; Panciera, T.; Gandin, A.; et al. The SWI/SNF complex is a mechanoregulated inhibitor of YAP and TAZ. Nature 2018, 563, 265–269. [Google Scholar] [CrossRef]
  109. Cho, Y.S.; Li, S.; Wang, X.; Zhu, J.; Zhuo, S.; Han, Y.; Yue, T.; Yang, Y.; Jiang, J. CDK7 regulates organ size and tumor growth by safeguarding the Hippo pathway effector Yki/Yap/Taz in the nucleus. Genes Dev. 2020, 34, 53–71. [Google Scholar] [CrossRef]
  110. Alhousami, T.; Diny, M.; Ali, F.; Shin, J.; Kumar, G.; Kumar, V.; Campbell, J.D.; Noonan, V.; Hanna, G.J.; Denis, G.V.; et al. Inhibition of LSD1 Attenuates Oral Cancer Development and Promotes Therapeutic Efficacy of Immune Checkpoint Blockade and YAP/TAZ Inhibition. Mol. Cancer Res. 2022, 20, 712–721. [Google Scholar] [CrossRef]
  111. Li, H.; Wu, B.K.; Kanchwala, M.; Cai, J.; Wang, L.; Xing, C.; Zheng, Y.; Pan, D. YAP/TAZ drives cell proliferation and tumour growth via a polyamine-eIF5A hypusination-LSD1 axis. Nat. Cell Biol. 2022, 24, 373–383. [Google Scholar] [CrossRef] [PubMed]
  112. 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. [Google Scholar] [CrossRef] [PubMed]
  113. Lee, B.S.; Park, D.I.; Lee, D.H.; Lee, J.E.; Yeo, M.K.; Park, Y.H.; Lim, D.S.; Choi, W.; Lee, D.H.; Yoo, G.; et al. Hippo effector YAP directly regulates the expression of PD-L1 transcripts in EGFR-TKI-resistant lung adenocarcinoma. Biochem. Biophys. Res. Commun. 2017, 491, 493–499. [Google Scholar] [CrossRef] [PubMed]
  114. Janse van Rensburg, H.J.; Azad, T.; Ling, M.; Hao, Y.; Snetsinger, B.; Khanal, P.; Minassian, L.M.; Graham, C.H.; Rauh, M.J.; Yang, X. The Hippo Pathway Component TAZ Promotes Immune Evasion in Human Cancer through PD-L1. Cancer Res. 2018, 78, 1457–1470. [Google Scholar] [CrossRef]
  115. Miao, J.; Hsu, P.C.; Yang, Y.L.; Xu, Z.; Dai, Y.; Wang, Y.; Chan, G.; Huang, Z.; Hu, B.; Li, H.; et al. YAP regulates PD-L1 expression in human NSCLC cells. Oncotarget 2017, 8, 114576–114587. [Google Scholar] [CrossRef]
  116. Hsu, P.C.; Miao, J.; Wang, Y.C.; Zhang, W.Q.; Yang, Y.L.; Wang, C.W.; Yang, C.T.; Huang, Z.; You, J.; Xu, Z.; et al. Inhibition of yes-associated protein down-regulates PD-L1 (CD274) expression in human malignant pleural mesothelioma. J. Cell Mol. Med. 2018, 22, 3139–3148. [Google Scholar] [CrossRef]
  117. Kim, M.H.; Kim, C.G.; Kim, S.K.; Shin, S.J.; Choe, E.A.; Park, S.H.; Shin, E.C.; Kim, J. YAP-Induced PD-L1 Expression Drives Immune Evasion in BRAFi-Resistant Melanoma. Cancer Immunol. Res. 2018, 6, 255–266. [Google Scholar] [CrossRef]
  118. Chen, N.; Golczer, G.; Ghose, S.; Lin, B.; Langenbucher, A.; Webb, J.; Bhanot, H.; Abt, N.B.; Lin, D.; Varvares, M.; et al. YAP1 maintains active chromatin state in head and neck squamous cell carcinomas that promotes tumorigenesis through cooperation with BRD4. Cell Rep. 2022, 39, 110970. [Google Scholar] [CrossRef]
  119. Moroishi, T.; Hayashi, T.; Pan, W.W.; Fujita, Y.; Holt, M.V.; Qin, J.; Carson, D.A.; Guan, K.L. The Hippo Pathway Kinases LATS1/2 Suppress Cancer Immunity. Cell 2016, 167, 1525–1539.e17. [Google Scholar] [CrossRef]
  120. Macleod, A.R. Abstract ND11: The discovery and characterization of ION-537: A next generation antisense oligonucleotide inhibitor of YAP1 in preclinical cancer models. Cancer Res. 2021, 81, ND11. [Google Scholar] [CrossRef]
  121. Lin, K.C.; Moroishi, T.; Meng, Z.; Jeong, H.S.; Plouffe, S.W.; Sekido, Y.; Han, J.; Park, H.W.; Guan, K.L. Regulation of Hippo pathway transcription factor TEAD by p38 MAPK-induced cytoplasmic translocation. Nat. Cell Biol. 2017, 19, 996–1002. [Google Scholar] [CrossRef] [PubMed]
  122. Jiao, S.; Wang, H.; Shi, Z.; Dong, A.; Zhang, W.; Song, X.; He, F.; Wang, Y.; Zhang, Z.; Wang, W.; et al. A peptide mimicking VGLL4 function acts as a YAP antagonist therapy against gastric cancer. Cancer Cell 2014, 25, 166–180. [Google Scholar] [CrossRef] [PubMed]
  123. Yuan, Y.; Park, J.; Feng, A.; Awasthi, P.; Wang, Z.; Chen, Q.; Iglesias-Bartolome, R. YAP1/TAZ-TEAD transcriptional networks maintain skin homeostasis by regulating cell proliferation and limiting KLF4 activity. Nat. Commun. 2020, 11, 1472. [Google Scholar] [CrossRef] [PubMed]
  124. 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. [Google Scholar] [CrossRef]
  125. Pobbati, A.V.; Han, X.; Hung, A.W.; Weiguang, S.; Huda, N.; Chen, G.Y.; Kang, C.; Chia, C.S.; Luo, X.; Hong, W.; et al. Targeting the Central Pocket in Human Transcription Factor TEAD as a Potential Cancer Therapeutic Strategy. Structure 2015, 23, 2076–2086. [Google Scholar] [CrossRef]
  126. Noland, C.L.; Gierke, S.; Schnier, P.D.; Murray, J.; Sandoval, W.N.; Sagolla, M.; Dey, A.; Hannoush, R.N.; Fairbrother, W.J.; Cunningham, C.N. Palmitoylation of TEAD Transcription Factors Is Required for Their Stability and Function in Hippo Pathway Signaling. Structure 2016, 24, 179–186. [Google Scholar] [CrossRef]
  127. 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. [Google Scholar] [CrossRef]
  128. Kim, N.G.; Gumbiner, B.M. Cell contact and Nf2/Merlin-dependent regulation of TEAD palmitoylation and activity. Proc. Natl. Acad. Sci. USA 2019, 116, 9877–9882. [Google Scholar] [CrossRef]
  129. Lu, W.; Wang, J.; Li, Y.; Tao, H.; Xiong, H.; Lian, F.; Gao, J.; Ma, H.; Lu, T.; Zhang, D.; et al. Discovery and biological evaluation of vinylsulfonamide derivatives as highly potent, covalent TEAD autopalmitoylation inhibitors. Eur. J. Med. Chem. 2019, 184, 111767. [Google Scholar] [CrossRef]
  130. Li, Q.; Sun, Y.; Jarugumilli, G.K.; Liu, S.; Dang, K.; Cotton, J.L.; Xiol, J.; Chan, P.Y.; DeRan, M.; Ma, L.; et al. Lats1/2 Sustain Intestinal Stem Cells and Wnt Activation through TEAD-Dependent and Independent Transcription. Cell Stem. Cell 2020, 26, 675–692.e8. [Google Scholar] [CrossRef]
  131. Holden, J.K.; Crawford, J.J.; Noland, C.L.; Schmidt, S.; Zbieg, J.R.; Lacap, J.A.; Zang, R.; Miller, G.M.; Zhang, Y.; Beroza, P.; et al. Small Molecule Dysregulation of TEAD Lipidation Induces a Dominant-Negative Inhibition of Hippo Pathway Signaling. Cell Rep. 2020, 31, 107809. [Google Scholar] [CrossRef] [PubMed]
  132. 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. [Google Scholar] [CrossRef] [PubMed]
  133. Kurppa, K.J.; Liu, Y.; To, C.; Zhang, T.; Fan, M.; Vajdi, A.; Knelson, E.H.; Xie, Y.; Lim, K.; Cejas, P.; et al. Treatment-Induced Tumor Dormancy through YAP-Mediated Transcriptional Reprogramming of the Apoptotic Pathway. Cancer Cell 2020, 37, 104–122.e12. [Google Scholar] [CrossRef] [PubMed]
  134. Tolcher, A.W.; Lakhani, N.J.; McKean, M.; Lingaraj, T.; Victor, L.; Sanchez-Martin, M.; Kacena, K.; Malek, K.S.; Santillana, S. A phase 1, first-in-human study of IK-930, an oral TEAD inhibitor targeting the Hippo pathway in subjects with advanced solid tumors. J. Clin. Oncol. 2022, 40, TPS3168. [Google Scholar] [CrossRef]
  135. Esposito, D.; Pant, I.; Shen, Y.; Qiao, R.F.; Yang, X.; Bai, Y.; Jin, J.; Poulikakos, P.I.; Aaronson, S.A. ROCK1 mechano-signaling dependency of human malignancies driven by TEAD/YAP activation. Nat. Commun. 2022, 13, 703. [Google Scholar] [CrossRef]
  136. 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. 2021, 27, 2326–2339. [Google Scholar] [CrossRef]
  137. Rosenbluh, J.; Nijhawan, D.; Cox, A.G.; Li, X.; Neal, J.T.; Schafer, E.J.; Zack, T.I.; Wang, X.; Tsherniak, A.; Schinzel, A.C.; et al. β-Catenin-driven cancers require a YAP1 transcriptional complex for survival and tumorigenesis. Cell 2012, 151, 1457–1473. [Google Scholar] [CrossRef]
  138. Taniguchi, K.; Wu, L.W.; Grivennikov, S.I.; de Jong, P.R.; Lian, I.; Yu, F.X.; Wang, K.; Ho, S.B.; Boland, B.S.; Chang, J.T.; et al. A gp130-Src-YAP module links inflammation to epithelial regeneration. Nature 2015, 519, 57–62. [Google Scholar] [CrossRef]
  139. Feng, X.; Arang, N.; Rigiracciolo, D.C.; Lee, J.S.; Yeerna, H.; Wang, Z.; Lubrano, S.; Kishore, A.; Pachter, J.A.; König, G.M.; et al. A Platform of Synthetic Lethal Gene Interaction Networks Reveals that the GNAQ Uveal Melanoma Oncogene Controls the Hippo Pathway through FAK. Cancer Cell 2019, 35, 457–472.e5. [Google Scholar] [CrossRef]
  140. Goto, Y.; Ando, T.; Izumi, H.; Feng, X.; Arang, N.; Gilardi, M.; Wang, Z.; Ando, K.; Gutkind, J.S. Muscarinic receptors promote castration-resistant growth of prostate cancer through a FAK-YAP signaling axis. Oncogene 2020, 39, 4014–4027. [Google Scholar] [CrossRef]
  141. Li, P.; Silvis, M.R.; Honaker, Y.; Lien, W.H.; Arron, S.T.; Vasioukhin, V. αE-catenin inhibits a Src-YAP1 oncogenic module that couples tyrosine kinases and the effector of Hippo signaling pathway. Genes Dev. 2016, 30, 798–811. [Google Scholar] [CrossRef] [PubMed]
  142. Mo, J.S.; Meng, Z.; Kim, Y.C.; Park, H.W.; Hansen, C.G.; Kim, S.; Lim, D.S.; Guan, K.L. Cellular energy stress induces AMPK-mediated regulation of YAP and the Hippo pathway. Nat. Cell Biol. 2015, 17, 500–510. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, W.; Xiao, Z.D.; Li, X.; Aziz, K.E.; Gan, B.; Johnson, R.L.; Chen, J. AMPK modulates Hippo pathway activity to regulate energy homeostasis. Nat. Cell Biol. 2015, 17, 490–499. [Google Scholar] [CrossRef] [PubMed]
  144. DeRan, M.; Yang, J.; Shen, C.H.; Peters, E.C.; Fitamant, J.; Chan, P.; Hsieh, M.; Zhu, S.; Asara, J.M.; Zheng, B.; et al. Energy stress regulates hippo-YAP signaling involving AMPK-mediated regulation of angiomotin-like 1 protein. Cell Rep. 2014, 9, 495–503. [Google Scholar] [CrossRef] [PubMed]
  145. Peng, C.; Zhu, Y.; Zhang, W.; Liao, Q.; Chen, Y.; Zhao, X.; Guo, Q.; Shen, P.; Zhen, B.; Qian, X.; et al. Regulation of the Hippo-YAP Pathway by Glucose Sensor O-GlcNAcylation. Mol. Cell 2017, 68, 591–604.e5. [Google Scholar] [CrossRef]
  146. Wu, X.; Yeerna, H.; Goto, Y.; Ando, T.; Wu, V.H.; Zhang, X.; Wang, Z.; Amornphimoltham, P.; Murphy, A.N.; Tamayo, P.; et al. Metformin Inhibits Progression of Head and Neck Squamous Cell Carcinoma by Acting Directly on Carcinoma-Initiating Cells. Cancer Res. 2019, 79, 4360–4370. [Google Scholar] [CrossRef]
  147. Gulati, S.; Desai, J.; Palackdharry, S.M.; Morris, J.C.; Zhu, Z.; Jandarov, R.; Riaz, M.K.; Takiar, V.; Mierzwa, M.; Gutkind, J.S.; et al. Phase 1 dose-finding study of metformin in combination with concurrent cisplatin and radiotherapy in patients with locally advanced head and neck squamous cell cancer. Cancer 2020, 126, 354–362. [Google Scholar] [CrossRef]
  148. Gutkind, J.S.; Molinolo, A.A.; Wu, X.; Wang, Z.; Nachmanson, D.; Harismendy, O.; Alexandrov, L.B.; Wuertz, B.R.; Ondrey, F.G.; Laronde, D.; et al. Inhibition of mTOR signaling and clinical activity of metformin in oral premalignant lesions. JCI Insight 2021, 6, e147096. [Google Scholar] [CrossRef]
  149. Vitale-Cross, L.; Molinolo, A.A.; Martin, D.; Younis, R.H.; Maruyama, T.; Patel, V.; Chen, W.; Schneider, A.; Gutkind, J.S. Metformin prevents the development of oral squamous cell carcinomas from carcinogen-induced premalignant lesions. Cancer Prev. Res. 2012, 5, 562–573. [Google Scholar] [CrossRef]
  150. Madera, D.; Vitale-Cross, L.; Martin, D.; Schneider, A.; Molinolo, A.A.; Gangane, N.; Carey, T.E.; McHugh, J.B.; Komarck, C.M.; Walline, H.M.; et al. Prevention of tumor growth driven by PIK3CA and HPV oncogenes by targeting mTOR signaling with metformin in oral squamous carcinomas expressing OCT3. Cancer Prev. Res. 2015, 8, 197–207. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Roles of the Hippo pathway and YAP/TAZ. TAOKs phosphorylate MAP4Ks and MST1/2. Activated MAP4Ks and MST1/2 phosphorylate LATS1/2. Activated LATS1/2 phosphorylates YAP/TAZ at multiple serine residues, leading to its cytoplasmic retention or degradation. Cell-cell interaction induces cytoplasmic sequestration of YAP/TAZ by AMOT and α-catenin. While the Hippo is inactive, dephosphorylated YAP/TAZ translocates to the nucleus, binds to TEAD, and enhances transcription of growth-related genes. Mechanical stress induces RhoA activation and F-actin formation, thereby inhibiting LATS1/2. Serum or LPA treatment inhibits MAP4K4 and MST1/2 through STRIPAK formation, which causes LATS1/2 inactivation. Heat stress induces LATS1/2 dephosphorylation and degradation. Abbreviations are: ras homolog family member A (RhoA), Rho GTPase binding protein 1 (RHPN1), Neurofibromin 2 (NF2; also known as Merlin), kidney and brain protein (KIBRA; also known as WWC1), STRN-interacting protein (STRIP1/2), sarcolemmal membrane-associated protein (SLMAP), the angiomotin (AMOT) family, Mitogen-activated protein kinase kinase kinase kinase 4 (MAP4K4), TAO kinases (TAOKs), mammalian STE20-like kinase 1 and 2 (MST1/2; MST2 is also known as STK3), salvador homolog 1 (SAV1), large tumor suppressor 1 and 2 (LATS1/2), MOB kinase activator 1A and 1B (MOB1A/B), Yes-associated protein (YAP), and transcriptional co-activator with PDZ binding motif (TAZ; also known as WWTR1), ras homolog family member A (RhoA), vestigial-like proteins (VGLL), TEA domain family members (TEAD), connective growth factor (CTGF), and cysteine-rich angiogenic inducer 61 (CYR61), lysophosphatidic acid (LPA), and G protein-coupled receptor (GPCR).
Figure 1. Roles of the Hippo pathway and YAP/TAZ. TAOKs phosphorylate MAP4Ks and MST1/2. Activated MAP4Ks and MST1/2 phosphorylate LATS1/2. Activated LATS1/2 phosphorylates YAP/TAZ at multiple serine residues, leading to its cytoplasmic retention or degradation. Cell-cell interaction induces cytoplasmic sequestration of YAP/TAZ by AMOT and α-catenin. While the Hippo is inactive, dephosphorylated YAP/TAZ translocates to the nucleus, binds to TEAD, and enhances transcription of growth-related genes. Mechanical stress induces RhoA activation and F-actin formation, thereby inhibiting LATS1/2. Serum or LPA treatment inhibits MAP4K4 and MST1/2 through STRIPAK formation, which causes LATS1/2 inactivation. Heat stress induces LATS1/2 dephosphorylation and degradation. Abbreviations are: ras homolog family member A (RhoA), Rho GTPase binding protein 1 (RHPN1), Neurofibromin 2 (NF2; also known as Merlin), kidney and brain protein (KIBRA; also known as WWC1), STRN-interacting protein (STRIP1/2), sarcolemmal membrane-associated protein (SLMAP), the angiomotin (AMOT) family, Mitogen-activated protein kinase kinase kinase kinase 4 (MAP4K4), TAO kinases (TAOKs), mammalian STE20-like kinase 1 and 2 (MST1/2; MST2 is also known as STK3), salvador homolog 1 (SAV1), large tumor suppressor 1 and 2 (LATS1/2), MOB kinase activator 1A and 1B (MOB1A/B), Yes-associated protein (YAP), and transcriptional co-activator with PDZ binding motif (TAZ; also known as WWTR1), ras homolog family member A (RhoA), vestigial-like proteins (VGLL), TEA domain family members (TEAD), connective growth factor (CTGF), and cysteine-rich angiogenic inducer 61 (CYR61), lysophosphatidic acid (LPA), and G protein-coupled receptor (GPCR).
Jpm 12 01544 g001
Figure 2. Landscape of the genetic alterations dysregulating the Hippo pathway and YAP/TAZ in HNSCC. The landscape was created by c-BioPortal for Cancer Genomics (https://www.cbioportal.org accessed on 19 September 2022) by analyzing the data of the HNSCC TCGA, PanCancer Atlas (2015). Abbreviations are: Epidermal growth factor receptor (EGFR), FAT atypical cadherin 1 (FAT1), tissue inhibitor of metalloproteinase-1 (TIMP-1), vav guanine nucleotide exchange factor 2 (VAV2), protein tyrosine phosphatase non-receptor type 14 (PTPN14), AT-rich interaction domain 1A (ARID1A), bromodomain containing protein 4 (BRD4), Cyclin dependent kinase 7 (CDK7), and Lysine-specific demethylase 1 (LSD1).
Figure 2. Landscape of the genetic alterations dysregulating the Hippo pathway and YAP/TAZ in HNSCC. The landscape was created by c-BioPortal for Cancer Genomics (https://www.cbioportal.org accessed on 19 September 2022) by analyzing the data of the HNSCC TCGA, PanCancer Atlas (2015). Abbreviations are: Epidermal growth factor receptor (EGFR), FAT atypical cadherin 1 (FAT1), tissue inhibitor of metalloproteinase-1 (TIMP-1), vav guanine nucleotide exchange factor 2 (VAV2), protein tyrosine phosphatase non-receptor type 14 (PTPN14), AT-rich interaction domain 1A (ARID1A), bromodomain containing protein 4 (BRD4), Cyclin dependent kinase 7 (CDK7), and Lysine-specific demethylase 1 (LSD1).
Jpm 12 01544 g002
Figure 3. Schematic of the Hippo pathway and YAP/TAZ regulatory components in HNSCC. When the Hippo pathway is active, activated LATS1/2 phosphorylates YAP/TAZ at multiple serine residues, leading to its cytoplasmic retention or degradation. When Hippo is inactive, dephosphorylated YAP/TAZ translocates to the nucleus, binds to TEAD, and enhances transcription of growth-related genes. Genetic alterations including FAT1, EGFR, VAV2, TIMP-1, HPV (E7), ARID1A, BRD4, LSD1, AP-1, CDK7, and YAP1 are involved in aberrant YAP/TAZ activation. Abbreviations are: Rac family small GTPase 1 (RAC1), Ca2+/calmodulin-dependent protein kinase II (CAMK2), and scribble homolog (SCRIB).
Figure 3. Schematic of the Hippo pathway and YAP/TAZ regulatory components in HNSCC. When the Hippo pathway is active, activated LATS1/2 phosphorylates YAP/TAZ at multiple serine residues, leading to its cytoplasmic retention or degradation. When Hippo is inactive, dephosphorylated YAP/TAZ translocates to the nucleus, binds to TEAD, and enhances transcription of growth-related genes. Genetic alterations including FAT1, EGFR, VAV2, TIMP-1, HPV (E7), ARID1A, BRD4, LSD1, AP-1, CDK7, and YAP1 are involved in aberrant YAP/TAZ activation. Abbreviations are: Rac family small GTPase 1 (RAC1), Ca2+/calmodulin-dependent protein kinase II (CAMK2), and scribble homolog (SCRIB).
Jpm 12 01544 g003
Figure 4. Schematic of the role of Hippo pathway and YAP/TAZ in cetuximab resistance. Cetuximab treatment inactivates EGFR, which in turn reduces tyrosine phosphorylation of MOB1 and activated LATS1/2, resulting in transient YAP/TAZ inactivation. However, unknown mechanisms, including reactivation or overexpression of other receptor types of tyrosine kinases (RTKs), re-activate YAP/TAZ, leading to intrinsic or acquired resistance to cetuximab treatment. Combination therapy with cetuximab and a TEAD inhibitor may be beneficial for avoiding YAP/TAZ reactivation and tumor relapse. AREG denotes amphiregulin (also abbreviated AREG).
Figure 4. Schematic of the role of Hippo pathway and YAP/TAZ in cetuximab resistance. Cetuximab treatment inactivates EGFR, which in turn reduces tyrosine phosphorylation of MOB1 and activated LATS1/2, resulting in transient YAP/TAZ inactivation. However, unknown mechanisms, including reactivation or overexpression of other receptor types of tyrosine kinases (RTKs), re-activate YAP/TAZ, leading to intrinsic or acquired resistance to cetuximab treatment. Combination therapy with cetuximab and a TEAD inhibitor may be beneficial for avoiding YAP/TAZ reactivation and tumor relapse. AREG denotes amphiregulin (also abbreviated AREG).
Jpm 12 01544 g004
Figure 5. Schematic of therapeutic drugs targeting the Hippo pathway and YAP/TAZ in HNSCC. Metformin, Src, RhoA, Rac1, BET, LSD1, MEK, CDK7, and TEAD inhibitors are potential drugs for HNSCC treatment. AMPKa2 denotes AMP-activated catalytic subunit alpha 2 (also abbreviated AMPK).
Figure 5. Schematic of therapeutic drugs targeting the Hippo pathway and YAP/TAZ in HNSCC. Metformin, Src, RhoA, Rac1, BET, LSD1, MEK, CDK7, and TEAD inhibitors are potential drugs for HNSCC treatment. AMPKa2 denotes AMP-activated catalytic subunit alpha 2 (also abbreviated AMPK).
Jpm 12 01544 g005
Table 1. Representative potential drugs targeting the Hippo pathway undergoing clinical trials. Potential drugs undergoing clinical trials were searched using “head and neck cancer” or “solid, unspecified” terms using the Pharmaprojects intuitive interface of Citeline. Drugs shown as “Preclinical” were removed.
Table 1. Representative potential drugs targeting the Hippo pathway undergoing clinical trials. Potential drugs undergoing clinical trials were searched using “head and neck cancer” or “solid, unspecified” terms using the Pharmaprojects intuitive interface of Citeline. Drugs shown as “Preclinical” were removed.
Drug NameCompanyTargetClinical Trials
(ClinicalTrials.gov Identifier)
Cancer Type
IK-930Ikena OncologyTEADPhase I (NCT05228015)Solid, unspecified
ION-537Ionis PharmacueticalsYAPPhase I (NCT04659096)Solid, unspecified
BTX-A51BioTheryXCDK7Phase I (NCT04872166)Solid, unspecified
fadraciclibCyclacelCDK7Phase II (NCT04983810)Solid, unspecified
samuraciclibCarrick Therapeutics, EvotecCDK7Phase II (NCT03363893)Unspecified
SY-5609Syros PharmaceuticalsCDK7Phase I (NCT04247126)Solid, unspecified
XL-102ExelixisCDK7Phase I (NCT04726332)Solid, unspecified
trametinibNovartisMEKPhase II (NCT01376310)Solid, unspecified
binimetinibPfizerMEKPhase II (NCT01885195)Solid, unspecified
cobimetinibRocheMEKPhase II (NCT02639546)Solid, unspecified
mirdametinibSpringWorks TherapeuticsMEKPhase II (NCT05054374)Solid, unspecified
BI-894999Boehringer IngelheimBRD4Phase I (NCT02516553)Solid, unspecified
BPI-23314Betta PharmaceuticalsBRD4Phase I (CTR20192223, China FDA)Solid, unspecified
NUV-868Nuvation BioBRD4Phase II (NCT05252390)Solid, unspecified
PLX-2853Daiichi SankyoBRD4Phase II (NCT03297424)Solid, unspecified
SYHA-1801CSPC PharmaceuticalBRD4Phase I (NCT04309968)Solid, unspecified
JBI-802Jubilant Life SciencesLSD1Phase II (NCT05268666)Solid, unspecified
seclidemstatSalarius PharmaceuticalsLSD1Phase II (NCT05266196)Solid, unspecified
CC 90011Bristol-Myers SquibbLSD1Phase II (NCT02875223)Solid, unspecified
dasatinibBristol-Myers SquibbSrcPhase II (NCT00882583)Head and neck
repotrectinibTurning Point Therapeutics,
Zai Lab
SrcPhase III (NCT05004116)Solid, unspecified
VAL-201ValiRxSrcPhase II (NCT02280317)Solid, unspecified
TPX-0046Turning Point TherapeuticsSrcPhase II (NCT04161391)Solid, unspecified
elzovantinibTurning Point TherapeuticsSrcPhase II (NCT03993873)Solid, unspecified
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ando, T.; Okamoto, K.; Shintani, T.; Yanamoto, S.; Miyauchi, M.; Gutkind, J.S.; Kajiya, M. Integrating Genetic Alterations and the Hippo Pathway in Head and Neck Squamous Cell Carcinoma for Future Precision Medicine. J. Pers. Med. 2022, 12, 1544. https://doi.org/10.3390/jpm12101544

AMA Style

Ando T, Okamoto K, Shintani T, Yanamoto S, Miyauchi M, Gutkind JS, Kajiya M. Integrating Genetic Alterations and the Hippo Pathway in Head and Neck Squamous Cell Carcinoma for Future Precision Medicine. Journal of Personalized Medicine. 2022; 12(10):1544. https://doi.org/10.3390/jpm12101544

Chicago/Turabian Style

Ando, Toshinori, Kento Okamoto, Tomoaki Shintani, Souichi Yanamoto, Mutsumi Miyauchi, J. Silvio Gutkind, and Mikihito Kajiya. 2022. "Integrating Genetic Alterations and the Hippo Pathway in Head and Neck Squamous Cell Carcinoma for Future Precision Medicine" Journal of Personalized Medicine 12, no. 10: 1544. https://doi.org/10.3390/jpm12101544

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