Molecular targets of primary cilia defects in cancer (Review)

  • Authors:
    • Fengying Yin
    • Zihao Wei
    • Fangman Chen
    • Chuan Xin
    • Qianming Chen
  • View Affiliations

  • Published online on: July 4, 2022     https://doi.org/10.3892/ijo.2022.5388
  • Article Number: 98
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Primary cilia are hair‑like organelles that are present on the majority of mammalian cells. They are regarded as the regulatory ‘hub’ of cell functions due to their indispensable roles for several signaling pathways, such as Hh and Wnt pathways. Originally, cilia defects were found to cause a panoply of human diseases commonly referred to as ‘ciliopathies’. Evidence is accumulating that cilia defects are involved in the onset and development of cancer. Some proteins that cause cilia defects have been identified as oncogenes in multiple cancer types. Hence, understanding the pathways that cause cilia defects in cancer is of utmost importance for the development of novel cancer therapeutic targets. The present review article provides a critical overview of the molecular targets of primary cilia defects in cancer, and highlights their vast potential as therapeutic targets and novel biomarkers.

Introduction

Cilia are evolutionally conserved cell structures protruding from the cell surface, which are found ubiquitously across species from ancient protozoa to humans. They are usually divided into two categories: Motile cilia and primary cilia, both of which consist of the axoneme, matrix and ciliary membrane (Fig. 1). Unlike motile cilia, primary cilia usually lack the dynein motor proteins that power axonemal beating and are therefore immotile (1). For almost 100 years, cilia were considered as anomalous structures without any function, until they were found to be associated with the onset of multiple syndromes, such as Bardet-Biedl (2), Joubert (3), oral-facial-digital (OFD) (4), and Ellis-van Creveld syndrome (5).

There is abundant evidence to indicate that primary cilia tightly regulate numerous critical signaling pathways, such as the Hh, Wnt and Notch signaling pathways. Therefore, primary cilia are regarded as the regulatory ‘hub’ of cell functions (6,7). It is worth noting that, apart from basal cell carcinoma (BCC) (8) and medulloblastoma (9), in which tumorigenesis is cilia-dependent, cilia formation is compromised in multiple tumor types, including melanoma (10), pancreatic cancer (11), breast cancer (12), cholangiocarcinoma (CCA) (13), prostate cancer (14), renal cell carcinoma (RCC) (15) and oral squamous cell carcinoma (OSCC) (16). Since hundreds of proteins and numerous critical signaling pathways are regulated by primary cilia, their proper formation is critical for the integrity of signal transduction and various cellular processes (17). Therefore, molecules that cause primary cilia defects may provide novel approaches with which to attenuate tumor progression, which is the focus of the present review.

Primary cilia and the cell cycle

Uncontrolled cell proliferation and deregulation of the cell cycle are hallmarks of malignant tumor formation (18). This section mainly provides a description of the mechanisms through which ciliogenesis is closely associated with the cell cycle.

Cilia assembly and cell cycle

Ciliogenesis is an elaborately regulated process that begins when cells exit the mitotic cycle (Fig. 2). Ciliogenesis occurs through two main routes: The classic intracellular pathway, in which the short axoneme extends from the basal body before the latter reaches the cytoplasmic membrane (19); and the alternative pathway, in which the axoneme extends from the basal body only after the latter is anchored to the cytoplasmic membrane (20). In fibroblasts, cilia are assembled through the intracellular pathway. When the cell exits the cell cycle, cytoplasmic vesicles are anchored to the distal appendage of the mother centriole and then trigger the remodeling of the mother centriole into the basal body, which is the earliest ciliogenesis event that can be detected (21). The basal body-vesicle complex then migrates to the cell surface and is anchored to the plasma membrane. Under the transport of intraflagellar transport (IFT) complexes, proteins accumulate into cilia through the ‘ciliary gate’, and cilia gradually bulge from the cell surface. Conversely, in polarized epithelial and multiciliated cells, ciliogenesis occurs through the alternative pathway. The basal body localizes at the center of the apical membrane and interacts directly with the membrane, with cilia assembly occurring entirely in the plasma membrane [further details on the ciliogenic pathways observed have been previously described (22)].

Cilia disassembly and cell cycle

In contrast to cilia assembly, the mechanisms underlying cilia disassembly in normal or pathological conditions remain largely unknown (23). Following the stimulation of mitogens, the scaffolding protein human enhancer of filamentation 1 (HEF1) activates aurora kinase A (AURKA) in the basal body, which in turn phosphorylates histone deacetylase 6 (HDAC6) (24). Phosphorylated HDAC6 enters the cilia to remove the acetylation modification of axonemal microtubules, thus inducing cilia disassembly. However, the mechansims through which HEF1 is upregulated and recruited to the basal body following mitogen stimulation remain unclear. In addition to HEF1, Trichoplein (25) and Pitchfork (26) have also been proven to be activators of AURKA in the G1 and S phases, respectively.

In addition to HDAC6, kinesin family member (KIF)24 and KIF2a are also implicated in the de-polymerization of axonemal microtubules. The activity of KIF24 is enhanced by never-in-mitosis-A-related kinase 2 (NEK2) during the S and G2 phases (27), and KIF2a is activated by polo-like kinase 1 (PLK1) in the G2 and M phases (28). Notably, the regulatory mechanisms of AURKA, NEK2 and PLK1 involved in cilia disassembly are relatively conservative and in chronological order, ensuring the irreversibility of cilia resorption following cell cycle entry.

Since the centriole dually functions as the microtubule-organizing center during mitosis and as the basal body for ciliogenesis, a common hypothesis is that cilia anchor the centriole to the cell membrane, thereby depriving the cell of cycle entry (23). The abnormal presence of cilia has been shown to lead to cell cycle arrest observed in in vivo and in vitro models. Trichoplein or AURKA knockdown induce cilia to fail to disassemble, thus inducing G0/G1 arrest. More notably, this phenotype was reversed when cilia formation was prevented by the simultaneous knockdown of IFT20 (25). In a previous study, NudE neurodevelopment protein 1, localizing on the mother centriole, was found to be highly expressed during mitosis and then rapidly degraded when the cell became quiescent. Its silencing in zebrafish and cultured cells led to a significant increase in ciliary length and S phase arrest (29). In addition, in another study, Tctex1 deletion, mainly localizing on the transition zone and regulating ciliary resorption, was shown to resulted in cilia persistence and G1/S phase arrest (30). These studies suggest that cells may sense the presence of primary cilia to limit cell cycle entry; however, further research is required to prove this hypothesis.

Ciliogenesis is a cell cycle-regulated event in normal cells. However, few studies have closely implicated the ciliogenesis changes when the cell cycle is under the influence of malignancies. The most fundamental trait of malignant cells involves their ability to sustain proliferation (31). Therefore, a possible cause of cilia loss in malignant cells could be the increased proliferation. However, Menzl et al (12) found that the loss of primary cilia was not associated with an increased proliferative index (Ki67-positive cells) in breast cancer. Nobutani et al (32) hardly detected primary cilia in cell cycle-arrested human breast cancer cells. Moreover, in a study involving 110 patients with kidney cancer, reduced ciliary ratio was observed to be independent of cell proliferation (15). It is worth noting that ciliogenesis is not restrained in all malignant tumors. The ciliary ratio has been shown to be significantly increased in BCC tissues (33) and cilia have been identified in some subsets of human medulloblastomas which had activation in either Hh or Wnt signaling (9). These studies suggest that cilia defects are characteristic of oncogenic transformation, and this is not only due to the increased proliferation in malignancies.

Molecular targets that cause primary cilia defects in cancer

The abnormal expression of certain critical ciliary proteins, which have a number of the hallmarks of oncogenes, can result in severe cilia defects. Although only a few protein inhibitors have been used in clinical practice or in clinical trials (34), targeting these molecules still holds promise for cancer treatment (35). Nevertheless, a number of studies have been primarily observational and do not elaborate on the mechanisms of cancer containment by targeting these ciliary molecules. This section provides a summary and discussion of the mechanisms through which these molecules can cause cilia defects in cancer (Fig. 3).

Molecules regulating cilia assembly
Cell cycle-related kinase (CCRK)

CCRK plays a critical role in promoting G1 phase entry and is of interest as it is the closest vertebrate homolog to the long flagella 2, a critical protein controlling the length of flagella in chlamydomonas (36). CCRK localizes in the cytoplasm and tunes ciliary length and shape by coordinating the assembly of the ciliary axoneme and membrane (37,38). In addition, cell cycle regulation by CCRK has been found to be dependent on primary cilia. The mutation of CCRK in mice induces shortened and swollen cilia, which impairs the Hh pathway and leads to a constellation of developmental defects, including exencephaly, cleft palate and mild preaxial polydactyly/limb skeletal defects (39). Immature cilia with ultrastructural malformation are considered to be a true hallmark of glioblastoma tumors and to be associated with their unrestrained growth (40). Yang et al (41) found that the arrested ciliogenesis, and therefore tumor growth, were caused by the elevation of CCRK in glioblastoma. Furthermore, the depletion of CCRK markedly increased the ratio and length of primary cilia in glioblastoma cell lines, and notably, cells with restored cilia exhibited a block or severe delay in cell cycle progression, thus inhibiting tumor cell growth (41). It would be of interest to explore the mechanisms of action of CCRK in primary cilia in glioblastoma in vivo. On the whole, these finding indicate that the suppression of primary ciliogenesis by the upregulation of CCRK may be one of the mechanisms used by glioblastoma cells to provide a growth advantage.

PLK4

PLK4 is a conserved centrosomal protein (CEP) that plays a key role in the centriole duplication cycle in a concentration-dependent manner. The centrioles fail to duplicate when PLK4 malfunctions, while they overamplify when PLK4 is overexpressed (42). However, ciliogenesis is blocked, regardless of its deficiency or overexpression. Although abnormalities in centriole number and structure are commonly observed in cancer, cilia defects caused by PLK4 dysfunction in cancer were not reported until 2014. Shinmura et al (43) reported that the abnormally elevated expression of PLK4 was related to the absence of primary cilia in human gastric cancer. The upregulation of PLK4 mRNA expression was detected in half of the primary gastric cancers. Moreover, an in vitro experiment demonstrated that its overexpression directly caused centrosome amplification, leading to the suppression of ciliogenesis in gastric cancer cell lines (43). Another study overexpressed PLK4 in mice and found that the formation of primary cilia was prevented, cilia signaling was disrupted, and the formation of lymphomas and sarcomas in p53 null mice was advanced (44).

Thymosin β4 (Tβ4)

Tβ4, a 43-amino acid peptide (45), is widely distributed in several cell types, apart from red blood cells, and regulates actin cytoskeleton dynamics by sequestering globular actin monomers (46). Its elevation is highly associated with tumor malignancy in multiple cancer types, including melanoma, head and neck squamous cell carcinoma and bladder cancer (4749). The cilia ratio is increased by Tβ4 overexpression by upregulating the expression of nephronophthisis 3 (localizing to the basal body and indispensable for cilia integrity) in HeLa human cervical cancer cells (50). Further research by the same research group demonstrated that Tβ4-mediated cilia formation may control HeLa cell growth induced by di-(2-ethylhexyl) phthalate, a xenoestrogen with carcinogenic toxicity (51).

Thus far, the presence of cilia defects in cervical cancer remains contradictable and elusive. Primary cilia were rare in HeLa cells in the study by Alieva et al (52); however, in the study by Kowal and Falk (53), the ciliated population was significantly higher than previously anticipated. In addition, due to the extensive presence of primary cilia, the HeLa cell line was used as a model to explore cilia biology (54). However, these studies only examined primary cilia in the cultured cells and did not set up a strict control group of normal cervical epithelial cells. Therefore, further research is warranted at the tissue level.

Tectonic (TCTN)2

The TCTN family (including TCTN1, TCTN2 and TCTN3) forms complexes with proteins that localize to the transition zone of cilia, where they regulate the composition of the ciliary membrane in a tissue-dependent manner (55). Mutations in TCTNs lead to tissue-specific defects in cilia assembly and trafficking that underlie several ciliopathies (56). However, the available studies on the role of TCTNs in cancer are limited. TCTN1 elevation predicts poor clinical outcomes for patients with glioblastoma (57) and is indispensable for pancreatic cancer cell proliferation (58). Its association with primary cilia in these two cancer types has not yet been investigated, at least to the best of our knowledge. A previous study identified TCTN2 as an oncogene and a tumor marker in several cancer types, including colorectal, lung and ovarian cancer (59). Moreover, inhibiting the expression of TCTN2 has been found to significantly reduce colony formation and cell invasiveness, and impair the assembly of primary cilia in colon cancer cell lines. However, the existing studies are contradictory. Yasar et al (60) observed that the cilia number was elevated in colon adenocarcinoma compared with normal tissues, whereas the study by Rocha et al (61) demonstrated that the cilia number was decreased in colon cancer.

DAB2-interacting protein (DAB2IP)-KIF3a

DAB2IP, a member of the RasGTPase-activating protein family, has been identified as a tumor suppressor in several cancer types (62,63). Its loss is highly prevalent in different subtypes of RCC and is associated with a poor patient survival (64). A recent study found that DAB2IP knockdown impaired primary cilia formation by decreasing the expression of KIF3a (essential for cilia assembly and length maintenance) in kidney epithelial cells (65). In addition, the loss of KIF3a further promotes renal tumorigenesis, suggesting that primary cilia stability is part of the critical homeostatic machinery in renal epithelia (65).

Transforming acidic coiled-coil protein 3 (TACC3)

TACC3, characterized by a highly conserved C-terminal coil domain, is a key component of centrosome-microtubule dynamic networks (66). Recently, TACC3 has been identified as a potential prognostic marker and therapeutic target for various cancer types, such as breast (67) and lung cancer (68). Cilia ratio and length are gradually decreased during the progression from normal prostate to prostatic intraepithelial neoplasia and invasive prostate cancer, and that increase is accompanied by the activation of Wnt signaling (14). However, the mechanisms responsible for cilia defects in prostate cancer remain elusive. A recently published study demonstrated that TACC3 upregulation restrained ciliogenesis in prostate cancer cells (69). TACC3 can competitively bind filamin A and disrupt the michelin-filamin A interaction, which is necessary for centrosome migration to the apical membrane and cilia formation (70). Furthermore, TACC3 knockdown significantly restored the formation of primary cilia and inhibited tumorigenesis and tumor growth in vitro and in vivo, suggesting that targeting TACC3 may represent a novel approach for prostate cancer therapeutics (69).

Molecules regulating cilia disassembly
AURKA-HDAC6 signaling axis

As mentioned above, the AURKA-HDAC6 signaling axis plays a critical role in cilia disassembly. AURKA activation is common in multiple cancer types characterized by centrosomal amplification and genomic instability (71,72). The oncogenic AURKA appears to be the key node for the suppression of primary cilia in several types of cancer, including pancreatic ductal adenocarcinoma (PDAC) (73), clear cell RCC (ccRCC) (74), ovarian cancer (75), pheochromocytoma (76) and esophageal squamous cell carcinoma (ESCC) (77). In addition, in a previous study, the number of cilia was significantly decreased in AURKA-activated tissues from patients with OSCC (16). More importantly, AURKA inhibition has been shown to result in primary cilia re-expression and significantly inhibit tumor progression in PDAC, ccRCC and OSCC (16,73,78).

However, the target of AURKA for cilia absorption may vary between tumor types. Unlike ccRCC (74), pheochromocytoma (76) and ESCC (77), cilia loss induced by AURKA activation is independent of HDAC6 in PDAC (73), ovarian cancer (75) and OSCC (16). The specific target of AURKA in these cancer types is unclear. A previous study demonstrated that, in addition to HDAC6, AURKA also activated inositol polyphosphate 5-phosphatase E (INPP5E) (79), a ciliary protein whose absence results in ciliary destabilization and tumor progression in medulloblastoma (80). AURKA may promote cilia loss through INPP5E in PDAC, ovarian cancer and OSCC, which requires further verification.

Gradilone et al (81) focused on the mechanisms of HDAC6 on cilia resorption and cell proliferation in CCA. Their study reported a marked reduction in the number of cilia, which was accompanied by HDAC6 overexpression in clinical CCA tissues. Inhibiting HDAC6 by the pharmacologic inhibitor tubastatin-A or by shRNA could re-express primary cilia in CCA cell lines and decrease cell proliferation. More interestingly, the inhibitory effect of tubastatin-A was abolished after ciliogenesis was blocked by IFT88 knockdown, indicating that the inhibitory effect on cell proliferation through targeting HDAC6 is partially dependent on cilia restoration in CCA cells (81). The latest research from the same research group demonstrated that cilia disassembly in CCA was mediated by HDAC6-regulated autophagy in primary cilia (i.e., ciliophagy), suggesting that ciliophagy inhibition could be an important therapeutic target for CCA (82).

NEK2-KIF24 signaling axis

KIF24, localizing to the distal end of the centriole, depolymerizes the cilia microtubules and provokes cilia disassembly shortly following the stimulation of mitogens (27). The depolymerizing activity of KIF24 is enhanced by NEK2, a kinase only expressed in the S and G2 phases (83). The NEK2-KIF24 action at centriole prevents the aberrant assembly of cilia and keeps the de-ciliated state necessary for mitosis. This mechanism of inhibiting primary ciliogenesis is temporally distinct from the well-established AURKA-HDAC6 pathway by blocking the nucleation of cilia from the basal body (83).

NEK2 has been identified as an oncogene in various cancer types, including myeloma and breast cancer (8486). Cilia loss is considered to be a characteristic of breast cancer, and the mouse model further confirmed its importance in the development of breast cancer. Although the absence of primary cilia was not found to directly cause breast cancer in mice, it led to the earlier formation and accelerated the growth of cancer, and was associated with a higher grade and metastasis (87). Kim et al (83) found that NEK2-KIF24 was overexpressed in breast cancer cell lines, and their ablation resulted in the re-expression of primary cilia, thereby reducing the proliferation of cancer cells. However, NEK2 has not been reported to induce the loss of primary cilia in other tumors apart from breast cancer.

Tubulin tyrosine ligase like 3 (TTLL3)

The glutamylation of tubulin in mammals can be catalyzed by nine glutamate ligases, in which only two enzymes, TTLL3 and TTLL8, are capable of initiating glycylation on microtubules (88). Glycylation has been found to function as a critical regulator of ciliary disassembly in motile cilia and flagella, while previous studies that used zebrafish or mouse models proved that the integrity of primary cilia was also dependent on this post-translational modification on tubulin (89,90). TTLL3 is the only expressed glycylase in colon tissue, and its knockout in mice has been shown to not only cause a marked reduction in the number of cilia, but to also lead to a markedly increased cell proliferation rate in the colon epithelium. In addition, the lower expression level of TTLL3 has been shown to be significantly associated with the development of colorectal carcinoma, suggesting that cilia defects caused by TTLL3 may serve as a prognostic marker for this type cancer (61).

Molecules playing dual roles in cilia biology
Tuberous sclerosis complex (TSC)-mTOR complex 1 (mTORC1) pathway

TSC is a genetic syndrome with widespread dysplastic and multisystemic tumors, caused by the mutations in the tumor suppressor genes, TSC1 and TSC2 (91). TSC1 and TSC2 form a heterodimer that inhibits mTOR signaling by inactivating mTORC1, a type of complex sensitive to rapamycin (92). TSC1, localizing to the basal body (93), is frequently heterozygously lost in ccRCC (94). Activating mTOR signaling by silencing TSC1 lengthened primary cilia in zebrafish (95) and mouse (96) models, and inhibiting mTOR signaling by rapamycin, could return the ciliary length to normal (96). Rosengren et al (97) also found that mTORC1 inhibition by rapamycin resulted in shortened cilia. These studies appear to indicate that mTORC1 plays a positive role in ciliary maintenance. However, another study reported the negative effects of mTORC1 on primary cilia, in which ciliary length was increased following rapamycin treatment in a dose-dependent manner in renal epithelial and vascular endothelial cells (98).

Takahashi et al (99) found that mTORC1 played dual roles in ciliogenesis, since rapamycin treatment shortened the cilia length, whereas it promoted the cilia ratio in retinal pigmented epithelial (RPE1) cells. Furthermore, rapamycin treatment markedly restored the formation of primary cilia and attenuated cell proliferation in lung, kidney, breast and pancreatic cancer cell lines (99). Although the reasons for these discrepancies are unclear, in renal cancer (100) and patients with polycystic kidney disease (101), primary cilia defects are often accompanied by the aberrant activation of mTORC1. In addition, rapamycin and its derivatives have been widely used in the treatment of advanced renal (102) and breast cancer (103).

OFD1 and autophagy

The OFD1 gene was initially identified in OFD syndrome (104) and associated with other ciliopathies, such as Joubert syndrome and retinitis pigmentosa (105). As a component of the distal centriole, OFD1 builds centriole distal appendages, recruits IFT88 and stabilizes centriolar microtubules at a defined length to properly assemble cilia (106). Further research identified that OFD1, similar to CEP290 and pericentriolar material 1, is also the primary component of centriolar satellites, the particles surrounding centrosomes (107). Of note, OFD1 at the centriolar satellites functions as a suppressor of primary cilia, the opposite of the promotion of ciliogenesis observed in distal centrioles. Deficient autophagy-induced OFD1 populated centriolar satellites, leading to fewer and shorter primary cilia and, interestingly, partial OFD1 knockdown significantly restored the cilia formation in MCF7 cells, a human breast cancer cell line that completely lacks primary cilia (108).

The aforementioned study indicated a positive role of autophagy in the regulation of primary cilia formation, while another report recognized basal autophagy as a suppressor in ciliogenesis in primary Hürthle cell tumors. Cilia loss was caused by a high basal autophagic flux, and the inhibition of autophagosome formation notably restored the primary ciliogenesis in tumor cells (109). Pampliega et al (110) also found that autophagy inhibited ciliogenesis and cilia-associated signaling during normal nutritional conditions. On the other hand, Maharjan et al (111) reported that alterations in autophagy during serum-restimulation, irrespective of whether autophagy was activated or inhibited, prevented the disassembly of primary cilia in RPE1 cells. These studies suggest that the regulation of autophagy on primary cilia is likely dependent on the cellular context. Furthermore, the association between primary cilia and autophagy is bidirectional, since the abrogation of ciliogenesis partially inhibits autophagy (110). The regulatory mechanisms between autophagy and primary cilia are further complicated by other cellular events that are usually observed in cancer, such as the activation of AURKA, HDAC6 or other critical factors controlling cilia biology, and thus warrant further investigation (112).

Von Hippel-Lindau (VHL)

VHL has been widely accepted as a component of an E3 ubiquitin ligase that targets hypoxia-inducible factor α for ubiquitination and degradation in an oxygen-dependent manner (113). Its mutation results in VHL disease, the most well-known familial kidney cancer syndrome (114). ccRCC is the most common subtype of renal cancer, and is mainly sporadic. Of note, the VHL gene is inactivated in up to 87% of sporadic ccRCC cases (115). In addition, the re-expression of VHL protein is sufficient to suppress the formation of renal cancer in vivo, suggesting that VHL inactivation is a direct cause of renal tumorigenesis (116).

Primary cilia are almost absent in the renal cysts and ccRCC of patients with VHL (15). In addition, the histological manifestations of early-stage RCC, including increased disorganized cilia, occurred in the kidney of the VHL knockout zebrafish (117). However, in a previous study, ccRCC was not induced by a specific deletion of VHL in mouse renal epithelial cells, suggesting that additional mutations are required in mammals (118). The combined conditional inactivation of VHL and other tumor suppressor factors, Pten or tumor protein p53, gave rise to cilia ablation, renal cysts and neoplastic growth resembling human ccRCC (119,120). Of note, another study detected the mutations of ciliary genes in ~50% of 448 human ccRCC samples, suggesting that the dysfunction of primary cilia plays an important role in at least part of ccRCC (121).

The formation of primary cilia has been shown to be restored by re-expressing the wild-type VHL in VHL-defective ccRCC cell lines (122). However, the mechanisms responsible for cilia defects caused by VHL mutation remain elusive. Schermer et al (116) found that VHL localized on primary cilia and regulated the cilia maintenance by directing the growth of microtubules toward the cell periphery, which is a prerequisite for ciliogenesis. However, a different study revealed that VHL-knockdown led to cilia disassembly by upregulating the expression of NEK8, a cell cycle regulator (123). In addition, VHL knockdown increased AURKA expression by activating β-catenin, thus leading to cilia disassembly. Furthermore, the β-catenin responsive transcription inhibitor rescued cilia defects by inhibiting AURKA, opening new avenues for treatment with β-catenin inhibitors to rescue ciliogenesis in ccRCC (74).

Molecule regulating the transcription of ciliary genes
Split ends (Spen)

Spen, characterized by N-terminal RNA-binding motifs, is a large nuclear protein and a component of the HDAC corepressor complex (124). Spen regulates the expression of key transcriptional effectors in multiple signaling pathways (125) and has been established as a tumor suppressor gene by negatively regulating the transcription of estrogen receptor α targets in breast cancer (126). Recently, Spen was found to be co-expressed with the ciliogenic transcription factor, regulatory factor X family member 3. The knockdown of Spen considerably inhibited the formation of primary cilia, and its re-expression rescued ciliogenesis in breast cancer cells (127). Furthermore, the regulation of cell migration by Spen only occurred in those cells harboring primary cilia, indicating that Spen may coordinate cellular movement in a cilia-dependent manner in breast cancer (127).

Enhancer of zeste homolog 2 (EZH2)

EZH2, a histone methyltransferase, is part of polycomb repressive complex 2 (PRC2), which silences target genes epigenetically by catalyzing histone H3 tri-methylation (128). EZH2 is activated in a variety of cancer types and drives cancer progression by suppressing the expression of various tumor suppressor genes (129). Cilia loss is considered to be a promoter and a potential biomarker in melanoma development (10). A recent study (130) reported an inverse correlation between primary cilia and EZH2 expression levels during melanoma development. Activated EZH2 was a driver of melanoma oncogenesis by silencing genes related to ciliary integrity and thus deconstructing primary cilia. Cilia deconstruction further led to the activation of the Wnt/β-catenin pathway, a well-known oncogenic signaling pathway in melanoma. Strikingly, EZH2 activity blockage significantly induced primary ciliogenesis and cilia-dependent tumor growth-arrest, suggesting that rescuing ciliogenesis by targeting EZH2 may serve as a new strategy for melanoma treatment (130).

Syndesmos (SDOS)

SDOS was initially reported to co-localize with syndecan 4 cytoplasmic domain in focal contacts and promotes the assembly of focal adhesions (131). A recent study identified SDOS as a novel RNA-binding protein that interacted with TNF receptor-associated protein 1 at the endoplasmic reticulum to regulate mRNA translation (132). A small subset of mRNAs responsible for the primary ciliogenesis was regulated post-transcriptionally by SDOS, including transmembrane protein 67, coiled-coil and C2 domain containing 2A and Kif7, known as the ciliopathy-associated genes. Furthermore, the regulatory effect of SDOS on primary cilia was further proven in HeLa cells. The number and length of primary cilia were significantly increased in SDOS-silenced HeLa cells, whereas they were decreased in SDOS-overexpressing cells.

Others
KRAS proto-oncogene, GTPase (KRAS)

KRAS, having intrinsic GTPase activity, is a small GTP-binding protein and functions as a molecular switch for various cellular processes (133). Its mutation, a common driver in various cancers, locks the protein into the GTP-bound state and results in constitutive signaling, which gives a growth advantage to mutated cells, thereby leading to the development of cancer (134,135). In PDAC, KRAS is the most commonly mutated gene, which is present in >90% of tumor cells (136). In addition, the activation of the oncogenic KRAS allele directly induced the formation of PDAC in mice and blocked ciliogenesis in cancer cells (11). Of note, cilia defects were rescued by inhibiting KRAS effector pathways. The present study raises the possibility that aberrant KRAS signaling may promote carcinogenesis by inducing cilia loss in PDAC.

Conclusions and future perspectives

Primary cilia were previously considered as vestigial organelles, while recent advances have recognized the complex biological functions of these unique structures in diseases and cancer. Abnormal signaling pathways lead to uncontrolled proliferation (137), drug resistance (138) and immune escape in cancer (139). As the ‘hub’ for multiple signaling receptors and downstream effector molecules, primary cilia are considered to be the critical regulatory center for inducing pathway defects (140). Primary cilia are likely to impact cancer development in multiple ways, including affecting the cell cycle process (83), mediating signal transduction (141) and regulating the response to therapy targeted to ciliary proteins or related pathways (142). Therefore, several studies have highlighted the possible application of cilia dysfunction in the early diagnosis and prognosis pf cancer.

Primary cilia have been recognized as an important therapeutic target, although the mechanisms of cilia defects are variable in the context of each cancer type (Fig. 3). As already aforementioned, several key factors regulating primary cilia, such as AURKA, HDAC6 and NEK2, have been proven to function as critical oncogenes in the development of various cancer types (13,71,122). Targeting these oncogenic factors and rescuing ciliogenesis can restrain tumor growth, particularly when simultaneously targeting multiple proteins that cause cilia defects. Therefore, targeting these molecules to develop novel therapeutic approaches has been an emerging field in the research of pancreatic, lung, kidney and breast cancer (143).

Although a number of researchers have reported that tumor growth was arrested by re-expressing primary cilia in vivo and in vitro, further extensive research is warranted before targeting cilia can be used in cancer treatment. Firstly, the mechanisms through which primary cilia regulate carcinogenesis differs between cancer types and within cancer subtypes. Secondly, targeting cilia can both restrain tumor progression and induce kinase inhibitor resistance, which appears to prevent achieving good therapeutic effects (142). Primary cilia are complex organelles whose structure, arrangement and function are highly regulated. However, further studies are required to decipher the complex signals they transmit. Nevertheless, existing studies have suggested that the status of primary cilia should play a role in the decision making for precise and personalized treatment.

Acknowledgements

Not applicable.

Funding

The present study was supported by the Key Research and Development Program of Zhejiang Province (grant no. 2021C03074) and the Fundamental Research Funds for the Central Universities (grant no. K20220177).

Availability of data and materials

Not applicable.

Authors' contributions

FY performed the literature search and wrote the manuscript. ZW, CX and FC collected the relevant references and edited the manuscript. QC supervised and revised the manuscript. All authors have read and approved the final manuscript. Data authentication is not applicable.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Mitchison HM and Valente EM: Motile and non-motile cilia in human pathology: From function to phenotypes. J Pathol. 241:294–309. 2017. View Article : Google Scholar : PubMed/NCBI

2 

Ansley SJ, Badano JL, Blacque OE, Hill J, Hoskins BE, Leitch CC, Kim JC, Ross AJ, Eichers ER, Teslovich TM, et al: Basal body dysfunction is a likely cause of pleiotropic Bardet-Biedl syndrome. Nature. 425:628–633. 2003. View Article : Google Scholar : PubMed/NCBI

3 

Sattar S and Gleeson JG: The ciliopathies in neuronal development: A clinical approach to investigation of Joubert syndrome and Joubert syndrome-related disorders. Dev Med Child Neurol. 53:793–798. 2011. View Article : Google Scholar : PubMed/NCBI

4 

Franco B and Thauvin-Robinet C: Update on oral-facial-digital syndromes (OFDS). Cilia. 5:122016. View Article : Google Scholar : PubMed/NCBI

5 

Ruiz-Perez VL, Blair HJ, Rodriguez-Andres ME, Blanco MJ, Wilson A, Liu YN, Miles C, Peters H and Goodship JA: Evc is a positive mediator of Ihh-regulated bone growth that localises at the base of chondrocyte cilia. Development. 134:2903–2912. 2007. View Article : Google Scholar : PubMed/NCBI

6 

Singla V and Reiter JF: The primary cilium as the cell's antenna: Signaling at a sensory organelle. Science. 313:629–633. 2006. View Article : Google Scholar : PubMed/NCBI

7 

Pala R, Alomari N and Nauli SM: Primary cilium-dependent signaling mechanisms. Int J Mol Sci. 18:22722017. View Article : Google Scholar : PubMed/NCBI

8 

Wong SY, Seol AD, So PL, Ermilov AN, Bichakjian CK, Epstein EH Jr, Dlugosz AA and Reiter JF: Primary cilia can both mediate and suppress Hedgehog pathway-dependent tumorigenesis. Nat Med. 15:1055–1061. 2009. View Article : Google Scholar : PubMed/NCBI

9 

Han YG, Kim HJ, Dlugosz AA, Ellison DW, Gilbertson RJ and Alvarez-Buylla A: Dual and opposing roles of primary cilia in medulloblastoma development. Nat Med. 15:1062–1065. 2009. View Article : Google Scholar : PubMed/NCBI

10 

Kim J, Dabiri S and Seeley ES: Primary cilium depletion typifies cutaneous melanoma in situ and malignant melanoma. PLoS One. 6:e274102011. View Article : Google Scholar : PubMed/NCBI

11 

Seeley ES, Carriere C, Goetze T, Longnecker DS and Korc M: Pancreatic cancer and precursor pancreatic intraepithelial neoplasia lesions are devoid of primary cilia. Cancer Res. 69:422–430. 2009. View Article : Google Scholar : PubMed/NCBI

12 

Menzl I, Lebeau L, Pandey R, Hassounah NB, Li FW, Nagle R, Weihs K and McDermott KM: Loss of primary cilia occurs early in breast cancer development. Cilia. 3:72014. View Article : Google Scholar : PubMed/NCBI

13 

Mansini AP, Lorenzo Pisarello MJ, Thelen KM, Cruz-Reyes M, Peixoto E, Jin S, Howard BN, Trussoni CE, Gajdos GB, LaRusso NF, et al: MicroRNA (miR)-433 and miR-22 dysregulations induce histone-deacetylase-6 overexpression and ciliary loss in cholangiocarcinoma. Hepatology. 68:561–573. 2018. View Article : Google Scholar : PubMed/NCBI

14 

Hassounah NB, Nagle R, Saboda K, Roe DJ, Dalkin BL and McDermott KM: Primary cilia are lost in preinvasive and invasive prostate cancer. PLoS One. 8:e685212013. View Article : Google Scholar : PubMed/NCBI

15 

Basten SG, Willekers S, Vermaat JS, Slaats GG, Voest EE, van Diest PJ and Giles RH: Reduced cilia frequencies in human renal cell carcinomas versus neighboring parenchymal tissue. Cilia. 2:22013. View Article : Google Scholar : PubMed/NCBI

16 

Yin F, Chen Q, Shi Y, Xu H, Huang J, Qing M, Zhong L, Li J, Xie L and Zeng X: Activation of EGFR-Aurora A induces loss of primary cilia in oral squamous cell carcinoma. Oral Dis. 28:621–630. 2022. View Article : Google Scholar : PubMed/NCBI

17 

Yin F, Chen Q, Shi Y, Xu H, Huang J, Qing M, Zhong L, Li J, Xie L and Zeng X: Activation of EGFR-Aurora A induces loss of primary cilia in oral squamous cell carcinoma. Oral Dis. 28:621–630. 2022. View Article : Google Scholar : PubMed/NCBI

18 

Hanahan D and Weinberg RA: The hallmarks of cancer. Cell. 100:57–70. 2000. View Article : Google Scholar : PubMed/NCBI

19 

Sorokin S: Centrioles and the formation of rudimentary cilia by fibroblasts and smooth muscle cells. J Cell Biol. 15:363–377. 1962. View Article : Google Scholar : PubMed/NCBI

20 

Sorokin SP: Reconstructions of centriole formation and ciliogenesis in mammalian lungs. J Cell Sci. 3:207–230. 1968. View Article : Google Scholar : PubMed/NCBI

21 

Reiter JF, Blacque OE and Leroux MR: The base of the cilium: Roles for transition fibres and the transition zone in ciliary formation, maintenance and compartmentalization. EMBO Rep. 13:608–618. 2012. View Article : Google Scholar : PubMed/NCBI

22 

Bernabe-Rubio M and Alonso MA: Routes and machinery of primary cilium biogenesis. Cell Mol Life Sci. 74:4077–4095. 2017. View Article : Google Scholar : PubMed/NCBI

23 

Sanchez I and Dynlacht BD: Cilium assembly and disassembly. Nat Cell Biol. 18:711–717. 2016. View Article : Google Scholar : PubMed/NCBI

24 

Pugacheva EN, Jablonski SA, Hartman TR, Henske EP and Golemis EA: HEF1-dependent Aurora A activation induces disassembly of the primary cilium. Cell. 129:1351–1363. 2007. View Article : Google Scholar : PubMed/NCBI

25 

Inoko A, Matsuyama M, Goto H, Ohmuro-Matsuyama Y, Hayashi Y, Enomoto M, Ibi M, Urano T, Yonemura S, Kiyono T, et al: Trichoplein and Aurora A block aberrant primary cilia assembly in proliferating cells. J Cell Biol. 197:391–405. 2012. View Article : Google Scholar : PubMed/NCBI

26 

Kinzel D, Boldt K, Davis EE, Burtscher I, Trümbach D, Diplas B, Attié-Bitach T, Wurst W, Katsanis N, Ueffing M and Lickert H: Pitchfork regulates primary cilia disassembly and left-right asymmetry. Dev Cell. 19:66–77. 2010. View Article : Google Scholar : PubMed/NCBI

27 

Kobayashi T, Tsang WY, Li J, Lane W and Dynlacht BD: Centriolar kinesin Kif24 interacts with CP110 to remodel microtubules and regulate ciliogenesis. Cell. 145:914–925. 2011. View Article : Google Scholar : PubMed/NCBI

28 

Miyamoto T, Hosoba K, Ochiai H, Royba E, Izumi H, Sakuma T, Yamamoto T, Dynlacht BD and Matsuura S: The Microtubule-depolymerizing activity of a mitotic kinesin protein KIF2A drives primary cilia disassembly coupled with cell proliferation. Cell Rep. 10:664–673. 2015. View Article : Google Scholar : PubMed/NCBI

29 

Kim S, Zaghloul NA, Bubenshchikova E, Oh EC, Rankin S, Katsanis N, Obara T and Tsiokas L: Nde1-mediated inhibition of ciliogenesis affects cell cycle Re-entry. Nat Cell Biol. 13:351–360. 2011. View Article : Google Scholar : PubMed/NCBI

30 

Li A, Saito M, Chuang JZ, Tseng YY, Dedesma C, Tomizawa K, Kaitsuka T and Sung CH: Ciliary transition zone activation of phosphorylated Tctex-1 controls ciliary resorption, S-phase entry and fate of neural progenitors. Nat Cell Biol. 13:402–411. 2011. View Article : Google Scholar : PubMed/NCBI

31 

Hanahan D: Hallmarks of cancer: New dimensions. Cancer Discov. 12:31–46. 2022. View Article : Google Scholar : PubMed/NCBI

32 

Nobutani K, Shimono Y, Yoshida M, Mizutani K, Minami A, Kono S, Mukohara T, Yamasaki T, Itoh T, Takao S, et al: Absence of primary cilia in cell cycle-arrested human breast cancer cells. Genes Cells. 19:141–152. 2014. View Article : Google Scholar : PubMed/NCBI

33 

Yang N, Leung EL, Liu C, Li L, Eguether T, Jun Yao XJ, Jones EC, Norris DA, Liu A, Clark RA, et al: INTU is essential for oncogenic Hh signaling through regulating primary cilia formation in basal cell carcinoma. Oncogene. 36:4997–5005. 2017. View Article : Google Scholar : PubMed/NCBI

34 

Frett B, Brown RV, Ma M, Hu W, Han H and Li HY: Therapeutic melting pot of never in mitosis gene a related kinase 2 (Nek2): A perspective on Nek2 as an oncology target and recent advancements in Nek2 small molecule inhibition. J Med Chem. 57:5835–5844. 2014. View Article : Google Scholar : PubMed/NCBI

35 

Yang PH, Zhang L, Zhang YJ, Zhang J and Xu WF: HDAC6: Physiological function and its selective inhibitors for cancer treatment. Drug Discov Ther. 7:233–242. 2013. View Article : Google Scholar : PubMed/NCBI

36 

Tam LW, Wilson NF and Lefebvre PA: A CDK-related kinase regulates the length and assembly of flagella in Chlamydomonas. J Cell Biol. 176:819–829. 2007. View Article : Google Scholar : PubMed/NCBI

37 

Maurya AK, Rogers T and Sengupta P: A CCRK and a MAK kinase modulate cilia branching and length via regulation of axonemal microtubule dynamics in caenorhabditis elegans. Curr Biol. 29:1286–1300.e4. 2019. View Article : Google Scholar : PubMed/NCBI

38 

Ko HW, Norman RX, Tran J, Fuller KP, Fukuda M and Eggenschwiler JT: Broad-minded links cell cycle-related kinase to cilia assembly and hedgehog signal transduction. Dev Cell. 18:237–247. 2010. View Article : Google Scholar : PubMed/NCBI

39 

Snouffer A, Brown D, Lee H, Walsh J, Lupu F, Norman R, Lechtreck K, Ko HW and Eggenschwiler J: Cell Cycle-related kinase (CCRK) regulates ciliogenesis and Hedgehog signaling in mice. PLoS Genet. 13:e10069122017. View Article : Google Scholar : PubMed/NCBI

40 

Moser JJ, Fritzler MJ and Rattner JB: Ultrastructural characterization of primary cilia in pathologically characterized human glioblastoma multiforme (GBM) tumors. BMC Clin Pathol. 14:402014. View Article : Google Scholar : PubMed/NCBI

41 

Yang Y, Roine N and Makela TP: CCRK depletion inhibits glioblastoma cell proliferation in a cilium-dependent manner. EMBO Rep. 14:741–747. 2013. View Article : Google Scholar : PubMed/NCBI

42 

Hori A, Barnouin K, Snijders AP and Toda T: A non-canonical function of Plk4 in centriolar satellite integrity and ciliogenesis through PCM1 phosphorylation. EMBO Rep. 17:326–337. 2016. View Article : Google Scholar : PubMed/NCBI

43 

Shinmura K, Kurabe N, Goto M, Yamada H, Natsume H, Konno H and Sugimura H: PLK4 overexpression and its effect on centrosome regulation and chromosome stability in human gastric cancer. Mol Biol Rep. 41:6635–6644. 2014. View Article : Google Scholar : PubMed/NCBI

44 

Coelho PA, Bury L, Shahbazi MN, Liakath-Ali K, Tate PH, Wormald S, Hindley CJ, Huch M, Archer J, Skarnes WC, et al: Over-expression of Plk4 induces centrosome amplification, loss of primary cilia and associated tissue hyperplasia in the mouse. Open Biol. 5:1502092015. View Article : Google Scholar : PubMed/NCBI

45 

Goldstein AL, Hannappel E, Sosne G and Kleinman HK: Thymosin β4: A multi-functional regenerative peptide. Basic properties and clinical applications. Expert Opin Biol Ther. 12:37–51. 2012. View Article : Google Scholar : PubMed/NCBI

46 

Safer D, Elzinga M and Nachmias VT: Thymosin beta 4 and Fx, an actin-sequestering peptide, are indistinguishable. J Biol Chem. 266:4029–4032. 1991. View Article : Google Scholar : PubMed/NCBI

47 

Cha HJ, Jeong MJ and Kleinman HK: Role of thymosin beta4 in tumor metastasis and angiogenesis. J Natl Cancer Inst. 95:1674–1680. 2003. View Article : Google Scholar : PubMed/NCBI

48 

Wang ZY, Zhang W, Yang JJ, Song DK, Wei JX and Gao S: Association of thymosin beta4 expression with clinicopathological parameters and clinical outcomes of bladder cancer patients. Neoplasma. 63:991–998. 2016. View Article : Google Scholar : PubMed/NCBI

49 

Chi LH, Chang WM, Chang YC, Chan YC, Tai CC, Leung KW, Chen CL, Wu AT, Lai TC, Li YJ and Hsiao M: Global Proteomics-based identification and validation of thymosin beta-4 X-Linked as a prognostic marker for head and neck squamous cell carcinoma. Sci Rep. 7:90312017. View Article : Google Scholar : PubMed/NCBI

50 

Lee JW, Kim HS and Moon EY: Thymosin beta-4 is a novel regulator for primary cilium formation by nephronophthisis 3 in HeLa human cervical cancer cells. Sci Rep. 9:68492019. View Article : Google Scholar : PubMed/NCBI

51 

Lee JW, Thuy PX, Han HK and Moon EY: Di-(2-ethylhexyl) phthalate-induced tumor growth is regulated by primary cilium formation via the axis of H2O2 production-thymosin beta-4 gene expression. Int J Med Sci. 18:1247–1258. 2021. View Article : Google Scholar : PubMed/NCBI

52 

Alieva IB, Gorgidze LA, Komarova YA, Chernobelskaya OA and Vorobjev IA: Experimental model for studying the primary cilia in tissue culture cells. Membr Cell Biol. 12:895–905. 1999.PubMed/NCBI

53 

Kowal TJ and Falk MM: Primary cilia found on HeLa and other cancer cells. Cell Biol Int. 39:1341–1347. 2015. View Article : Google Scholar : PubMed/NCBI

54 

Vorobyeva AG and Saunders AJ: Amyloid-beta interrupts canonical Sonic hedgehog signaling by distorting primary cilia structure. Cilia. 7:52018. View Article : Google Scholar : PubMed/NCBI

55 

Garcia-Gonzalo FR, Corbit KC, Sirerol-Piquer MS, Ramaswami G, Otto EA, Noriega TR, Seol AD, Robinson JF, Bennett CL, Josifova DJ, et al: A transition zone complex regulates mammalian ciliogenesis and ciliary membrane composition. Nat Genet. 43:776–784. 2011. View Article : Google Scholar : PubMed/NCBI

56 

Sang L, Miller JJ, Corbit KC, Giles RH, Brauer MJ, Otto EA, Baye LM, Wen X, Scales SJ, Kwong M, et al: Mapping the NPHP-JBTS-MKS protein network reveals ciliopathy disease genes and pathways. Cell. 145:513–528. 2011. View Article : Google Scholar : PubMed/NCBI

57 

Meng D, Chen Y, Zhao Y, Wang J, Yun D, Yang S, Chen J, Chen H and Lu D: Expression and prognostic significance of TCTN1 in human glioblastoma. J Transl Med. 12:2882014. View Article : Google Scholar : PubMed/NCBI

58 

Zhao S, Chen X, Wan M, Jiang X, Li C, Cui Y and Kang P: Tectonic 1 is a key regulator of cell proliferation in pancreatic cancer. Cancer Biother Radiopharm. 31:7–13. 2016. View Article : Google Scholar : PubMed/NCBI

59 

Cano-Rodriguez D, Campagnoli S, Grandi A, Parri M, Camilli E, Song C, Jin B, Lacombe A, Pierleoni A, Bombaci M, et al: TCTN2: A novel tumor marker with oncogenic properties. Oncotarget. 8:95256–95269. 2017. View Article : Google Scholar : PubMed/NCBI

60 

Yasar B, Linton K, Slater C and Byers R: Primary cilia are increased in number and demonstrate structural abnormalities in human cancer. J Clin Pathol. 70:571–574. 2017. View Article : Google Scholar : PubMed/NCBI

61 

Rocha C, Papon L, Cacheux W, Marques Sousa P, Lascano V, Tort O, Giordano T, Vacher S, Lemmers B, Mariani P, et al: Tubulin glycylases are required for primary cilia, control of cell proliferation and tumor development in colon. EMBO J. 33:2247–2260. 2014. View Article : Google Scholar : PubMed/NCBI

62 

Wang Z, Tseng CP, Pong RC, Chen H, McConnell JD, Navone N and Hsieh JT: The mechanism of growth-inhibitory effect of DOC-2/DAB2 in prostate cancer. Characterization of a novel GTPase-activating protein associated with N-terminal domain of DOC-2/DAB2. J Biol Chem. 277:12622–12631. 2002. View Article : Google Scholar : PubMed/NCBI

63 

Shen YJ, Kong ZL, Wan FN, Wang HK, Bian XJ, Gan HL, Wang CF and Ye DW: Downregulation of DAB2IP results in cell proliferation and invasion and contributes to unfavorable outcomes in bladder cancer. Cancer Sci. 105:704–712. 2014. View Article : Google Scholar : PubMed/NCBI

64 

Wang ZR, Wei JH, Zhou JC, Haddad A, Zhao LY, Kapur P, Wu KJ, Wang B, Yu YH, Liao B, et al: Validation of DAB2IP methylation and its relative significance in predicting outcome in renal cell carcinoma. Oncotarget. 7:31508–31519. 2016. View Article : Google Scholar : PubMed/NCBI

65 

Lin CJ, Dang A, Hernandez E and Hsieh JT: DAB2IP modulates primary cilia formation associated with renal tumorigenesis. Neoplasia. 23:169–180. 2021. View Article : Google Scholar : PubMed/NCBI

66 

Schneider L, Essmann F, Kletke A, Rio P, Hanenberg H, Wetzel W, Schulze-Osthoff K, Nürnberg B and Piekorz RP: The transforming acidic coiled coil 3 protein is essential for spindle-dependent chromosome alignment and mitotic survival. J Biol Chem. 282:29273–29283. 2007. View Article : Google Scholar : PubMed/NCBI

67 

Campo L and Breuer EK: Inhibition of TACC3 by a small molecule inhibitor in breast cancer. Biochem Biophys Res Commun. 498:1085–1092. 2018. View Article : Google Scholar : PubMed/NCBI

68 

Jiang F, Kuang B, Que Y, Lin Z, Yuan L, Xiao W, Peng R and Zhang X and Zhang X: The clinical significance of transforming acidic coiled-coil protein 3 expression in non-small cell lung cancer. Oncol Rep. 35:436–446. 2016. View Article : Google Scholar : PubMed/NCBI

69 

Qie Y, Wang L, Du E, Chen S, Lu C, Ding N, Yang K and Xu Y: TACC3 promotes prostate cancer cell proliferation and restrains primary cilium formation. Exp Cell Res. 390:1119522020. View Article : Google Scholar : PubMed/NCBI

70 

Adams M, Simms RJ, Abdelhamed Z, Dawe HR, Szymanska K, Logan CV, Wheway G, Pitt E, Gull K, Knowles MA, et al: A meckelin-filamin A interaction mediates ciliogenesis. Hum Mol Genet. 21:1272–1286. 2012. View Article : Google Scholar : PubMed/NCBI

71 

Goepfert TM, Adigun YE, Zhong L, Gay J, Medina D and Brinkley WR: Centrosome amplification and overexpression of aurora A are early events in rat mammary carcinogenesis. Cancer Res. 62:4115–4122. 2002.PubMed/NCBI

72 

Gritsko TM, Coppola D, Paciga JE, Yang L, Sun M, Shelley SA, Fiorica JV, Nicosia SV and Cheng JQ: Activation and overexpression of centrosome kinase BTAK/Aurora-A in human ovarian cancer. Clin Cancer Res. 9:1420–1426. 2003.PubMed/NCBI

73 

Kobayashi T, Nakazono K, Tokuda M, Mashima Y, Dynlacht BD and Itoh H: HDAC2 promotes loss of primary cilia in pancreatic ductal adenocarcinoma. EMBO Rep. 18:334–343. 2017. View Article : Google Scholar : PubMed/NCBI

74 

Dere R, Perkins AL, Bawa-Khalfe T, Jonasch D and Walker CL: β-catenin links von Hippel-Lindau to aurora kinase A and loss of primary cilia in renal cell carcinoma. J Am Soc Nephrol. 26:553–564. 2015. View Article : Google Scholar : PubMed/NCBI

75 

Egeberg DL, Lethan M, Manguso R, Schneider L, Awan A, Jørgensen TS, Byskov AG, Pedersen LB and Christensen ST: Primary cilia and aberrant cell signaling in epithelial ovarian cancer. Cilia. 1:152012. View Article : Google Scholar : PubMed/NCBI

76 

O'Toole SM, Watson DS, Novoselova TV, Romano LEL, King PJ, Bradshaw TY, Thompson CL, Knight MM, Sharp TV, Barnes MR, et al: Oncometabolite induced primary cilia loss in pheochromocytoma. Endocr Relat Cancer. 26:165–180. 2019. View Article : Google Scholar : PubMed/NCBI

77 

Chen Q, Li J, Yang X, Ma J, Gong F and Liu Y: Prdx1 promotes the loss of primary cilia in esophageal squamous cell carcinoma. BMC Cancer. 20:3722020. View Article : Google Scholar : PubMed/NCBI

78 

Xu J, Li H, Wang B, Xu Y, Yang J, Zhang X, Harten SK, Shukla D, Maxwell PH, Pei D and Esteban MA: VHL inactivation induces HEF1 and Aurora kinase A. J Am Soc Nephrol. 21:2041–2046. 2010. View Article : Google Scholar : PubMed/NCBI

79 

Plotnikova OV, Seo S, Cottle DL, Conduit S, Hakim S, Dyson JM, Mitchell CA and Smyth IM: INPP5E interacts with AURKA, linking phosphoinositide signaling to primary cilium stability. J Cell Sci. 128:364–372. 2015.PubMed/NCBI

80 

Conduit SE, Ramaswamy V, Remke M, Watkins DN, Wainwright BJ, Taylor MD, Mitchell CA and Dyson JM: A compartmentalized phosphoinositide signaling axis at cilia is regulated by INPP5E to maintain cilia and promote Sonic Hedgehog medulloblastoma. Oncogene. 36:5969–5984. 2017. View Article : Google Scholar : PubMed/NCBI

81 

Gradilone SA, Radtke BN, Bogert PS, Huang BQ, Gajdos GB and LaRusso NF: HDAC6 inhibition restores ciliary expression and decreases tumor growth. Cancer Res. 73:2259–2270. 2013. View Article : Google Scholar : PubMed/NCBI

82 

Peixoto E, Jin S, Thelen K, Biswas A, Richard S, Morleo M, Mansini A, Holtorf S, Carbone F, Pastore N, et al: HDAC6-dependent ciliophagy is involved in ciliary loss and cholangiocarcinoma growth in human cells and murine models. Am J Physiol Gastrointest Liver Physiol. 318:G1022–G1033. 2020. View Article : Google Scholar : PubMed/NCBI

83 

Kim S, Lee K, Choi JH, Ringstad N and Dynlacht BD: Nek2 activation of Kif24 ensures cilium disassembly during the cell cycle. Nat Commun. 6:80872015. View Article : Google Scholar : PubMed/NCBI

84 

Cappello P, Blaser H, Gorrini C, Lin DC, Elia AJ, Wakeham A, Haider S, Boutros PC, Mason JM, Miller NA, et al: Role of Nek2 on centrosome duplication and aneuploidy in breast cancer cells. Oncogene. 33:2375–2384. 2014. View Article : Google Scholar : PubMed/NCBI

85 

Zhou W, Yang Y, Xia J, Wang H, Salama ME, Xiong W, Xu H, Shetty S, Chen T, Zeng Z, et al: NEK2 induces drug resistance mainly through activation of efflux drug pumps and is associated with poor prognosis in myeloma and other cancers. Cancer Cell. 23:48–62. 2013. View Article : Google Scholar : PubMed/NCBI

86 

Hayward DG, Clarke RB, Faragher AJ, Pillai MR, Hagan IM and Fry AM: The centrosomal kinase Nek2 displays elevated levels of protein expression in human breast cancer. Cancer Res. 64:7370–7376. 2004. View Article : Google Scholar : PubMed/NCBI

87 

Hassounah NB, Nunez M, Fordyce C, Roe D, Nagle R, Bunch T and McDermott KM: Inhibition of ciliogenesis promotes hedgehog signaling, tumorigenesis, and metastasis in breast cancer. Mol Cancer Res. 15:1421–1430. 2017. View Article : Google Scholar : PubMed/NCBI

88 

Rogowski K, Juge F, van Dijk J, Wloga D, Strub JM, Levilliers N, Thomas D, Bré MH, Van Dorsselaer A, Gaertig J and Janke C: Evolutionary divergence of enzymatic mechanisms for posttranslational polyglycylation. Cell. 137:1076–1087. 2009. View Article : Google Scholar : PubMed/NCBI

89 

Bosch Grau M, Masson C, Gadadhar S, Rocha C, Tort O, Marques Sousa P, Vacher S, Bieche I and Janke C: Alterations in the balance of tubulin glycylation and glutamylation in photoreceptors leads to retinal degeneration. J Cell Sci. 130:938–949. 2017.PubMed/NCBI

90 

Pathak N, Austin CA and Drummond IA: Tubulin tyrosine ligase-like genes ttll3 and ttll6 maintain zebrafish cilia structure and motility. J Biol Chem. 286:11685–11695. 2011. View Article : Google Scholar : PubMed/NCBI

91 

Curatolo P, Bombardieri R and Jozwiak S: Tuberous sclerosis. Lancet. 372:657–668. 2008. View Article : Google Scholar : PubMed/NCBI

92 

Jozwiak J: Hamartin and tuberin: Working together for tumour suppression. Int J Cancer. 118:1–5. 2006. View Article : Google Scholar : PubMed/NCBI

93 

Hartman TR, Liu D, Zilfou JT, Robb V, Morrison T, Watnick T and Henske EP: The tuberous sclerosis proteins regulate formation of the primary cilium via a rapamycin-insensitive and polycystin 1-independent pathway. Hum Mol Genet. 18:151–163. 2009. View Article : Google Scholar : PubMed/NCBI

94 

Wilson C, Bonnet C, Guy C, Idziaszczyk S, Colley J, Humphreys V, Maynard J, Sampson JR and Cheadle JP: Tsc1 haploinsufficiency without mammalian target of rapamycin activation is sufficient for renal cyst formation in Tsc1+/- mice. Cancer Res. 66:7934–7938. 2006. View Article : Google Scholar : PubMed/NCBI

95 

DiBella LM, Park A and Sun Z: Zebrafish Tsc1 reveals functional interactions between the cilium and the TOR pathway. Hum Mol Genet. 18:595–606. 2009. View Article : Google Scholar : PubMed/NCBI

96 

Armour EA, Carson RP and Ess KC: Cystogenesis and elongated primary cilia in Tsc1-deficient distal convoluted tubules. Am J Physiol Renal Physiol. 303:F584–F592. 2012. View Article : Google Scholar : PubMed/NCBI

97 

Rosengren T, Larsen LJ, Pedersen LB, Christensen ST and Moller LB: TSC1 and TSC2 regulate cilia length and canonical Hedgehog signaling via different mechanisms. Cell Mol Life Sci. 75:2663–2680. 2018. View Article : Google Scholar : PubMed/NCBI

98 

Sherpa RT, Atkinson KF, Ferreira VP and Nauli SM: Rapamycin increases length and mechanosensory function of primary cilia in renal epithelial and vascular endothelial cells. Int Educ Res J. 2:91–97. 2016.PubMed/NCBI

99 

Takahashi K, Nagai T, Chiba S, Nakayama K and Mizuno K: Glucose deprivation induces primary cilium formation through mTORC1 inactivation. J Cell Sci. 131:jcs2087692018.PubMed/NCBI

100 

Huber TB, Walz G and Kuehn EW: mTOR and rapamycin in the kidney: Signaling and therapeutic implications beyond immunosuppression. Kidney Int. 79:502–511. 2011. View Article : Google Scholar : PubMed/NCBI

101 

Shillingford JM, Murcia NS, Larson CH, Low SH, Hedgepeth R, Brown N, Flask CA, Novick AC, Goldfarb DA, Kramer-Zucker A, et al: The mTOR pathway is regulated by polycystin-1, and its inhibition reverses renal cystogenesis in polycystic kidney disease. Proc Natl Acad Sci USA. 103:5466–5471. 2006. View Article : Google Scholar : PubMed/NCBI

102 

Zhang T and George DJ: Immunotherapy and targeted-therapy combinations mark a new era of kidney cancer treatment. Nat Med. 27:586–588. 2021. View Article : Google Scholar : PubMed/NCBI

103 

Fan Y, Sun T, Shao Z, Zhang Q, Ouyang Q, Tong Z, Wang S, Luo Y, Teng Y, Wang X, et al: Effectiveness of adding everolimus to the First-line treatment of advanced breast cancer in premenopausal women who experienced disease progression while receiving selective estrogen receptor modulators: A phase 2 randomized clinical trial. JAMA Oncol. 7:e2134282021. View Article : Google Scholar : PubMed/NCBI

104 

Ferrante MI, Giorgio G, Feather SA, Bulfone A, Wright V, Ghiani M, Selicorni A, Gammaro L, Scolari F, Woolf AS, et al: Identification of the gene for oral-facial-digital type I syndrome. Am J Hum Genet. 68:569–576. 2001. View Article : Google Scholar : PubMed/NCBI

105 

Wang J, Chen X, Wang F, Zhang J, Li P, Li Z, Xu J, Gao F, Jin C, Tian H, et al: OFD1, as a ciliary protein, exhibits neuroprotective function in photoreceptor degeneration models. PLoS One. 11:e01558602016. View Article : Google Scholar : PubMed/NCBI

106 

Singla V, Romaguera-Ros M, Garcia-Verdugo JM and Reiter JF: Ofd1, a human disease gene, regulates the length and distal structure of centrioles. Dev Cell. 18:410–424. 2010. View Article : Google Scholar : PubMed/NCBI

107 

Lopes CA, Prosser SL, Romio L, Hirst RA, O'Callaghan C, Woolf AS and Fry AM: Centriolar satellites are assembly points for proteins implicated in human ciliopathies, including oral-facial-digital syndrome 1. J Cell Sci. 124:600–612. 2011. View Article : Google Scholar : PubMed/NCBI

108 

Tang Z, Lin MG, Stowe TR, Chen S, Zhu M, Stearns T, Franco B and Zhong Q: Autophagy promotes primary ciliogenesis by removing OFD1 from centriolar satellites. Nature. 502:254–257. 2013. View Article : Google Scholar : PubMed/NCBI

109 

Lee J, Yi S, Kang YE, Chang JY, Kim JT, Sul HJ, Kim JO, Kim JM, Kim J, Porcelli AM, et al: Defective ciliogenesis in thyroid hurthle cell tumors is associated with increased autophagy. Oncotarget. 7:79117–79130. 2016. View Article : Google Scholar : PubMed/NCBI

110 

Pampliega O, Orhon I, Patel B, Sridhar S, Díaz-Carretero A, Beau I, Codogno P, Satir BH, Satir P and Cuervo AM: Functional interaction between autophagy and ciliogenesis. Nature. 502:194–200. 2013. View Article : Google Scholar : PubMed/NCBI

111 

Maharjan Y, Lee JN, Kwak S, Lim H, Dutta RK, Liu ZQ, So HS and Park R: Autophagy alteration prevents primary cilium disassembly in RPE1 cells. Biochem Biophys Res Commun. 500:242–248. 2018. View Article : Google Scholar : PubMed/NCBI

112 

Ko JY, Lee EJ and Park JH: Interplay between primary cilia and autophagy and its controversial roles in cancer. Biomol Ther (Seoul). 27:337–341. 2019. View Article : Google Scholar : PubMed/NCBI

113 

Kaelin WG Jr and Ratcliffe PJ: Oxygen sensing by metazoans: The central role of the HIF hydroxylase pathway. Mol Cell. 30:393–402. 2008. View Article : Google Scholar : PubMed/NCBI

114 

Maher ER and Kaelin WG Jr: von Hippel-Lindau disease. Medicine (Baltimore). 76:381–391. 1997. View Article : Google Scholar : PubMed/NCBI

115 

Arjumand W and Sultana S: Role of VHL gene mutation in human renal cell carcinoma. Tumour Biol. 33:9–16. 2012. View Article : Google Scholar : PubMed/NCBI

116 

Schermer B, Ghenoiu C, Bartram M, Müller RU, Kotsis F, Höhne M, Kühn W, Rapka M, Nitschke R, Zentgraf H, et al: The von Hippel-Lindau tumor suppressor protein controls ciliogenesis by orienting microtubule growth. J Cell Biol. 175:547–554. 2006. View Article : Google Scholar : PubMed/NCBI

117 

Noonan HR, Metelo AM, Kamei CN, Peterson RT, Drummond IA and Iliopoulos O: Loss of vhl in the zebrafish pronephros recapitulates early stages of human clear cell renal cell carcinoma. Dis Model Mech. 9:873–884. 2016. View Article : Google Scholar : PubMed/NCBI

118 

Frew IJ and Moch H: A clearer view of the molecular complexity of clear cell renal cell carcinoma. Annu Rev Pathol. 10:263–289. 2015. View Article : Google Scholar : PubMed/NCBI

119 

Albers J, Rajski M, Schonenberger D, Harlander S, Schraml P, von Teichman A, Georgiev S, Wild PJ, Moch H, Krek W and Frew IJ: Combined mutation of Vhl and Trp53 causes renal cysts and tumours in mice. EMBO Mol Med. 5:949–964. 2013. View Article : Google Scholar : PubMed/NCBI

120 

Frew IJ, Thoma CR, Georgiev S, Minola A, Hitz M, Montani M, Moch H and Krek W: pVHL and PTEN tumour suppressor proteins cooperatively suppress kidney cyst formation. EMBO J. 27:1747–1757. 2008. View Article : Google Scholar : PubMed/NCBI

121 

Harlander S, Schonenberger D, Toussaint NC, Prummer M, Catalano A, Brandt L, Moch H, Wild PJ and Frew IJ: Combined mutation in Vhl, Trp53 and Rb1 causes clear cell renal cell carcinoma in mice. Nat Med. 23:869–877. 2017. View Article : Google Scholar : PubMed/NCBI

122 

Esteban MA, Harten SK, Tran MG and Maxwell PH: Formation of primary cilia in the renal epithelium is regulated by the von Hippel-Lindau tumor suppressor protein. J Am Soc Nephrol. 17:1801–1806. 2006. View Article : Google Scholar : PubMed/NCBI

123 

Ding XF, Zhou J, Hu QY, Liu SC and Chen G: The tumor suppressor pVHL down-regulates never-in-mitosis A-related kinase 8 via hypoxia-inducible factors to maintain cilia in human renal cancer cells. J Biol Chem. 290:1389–1394. 2015. View Article : Google Scholar : PubMed/NCBI

124 

Oswald F, Kostezka U, Astrahantseff K, Bourteele S, Dillinger K, Zechner U, Ludwig L, Wilda M, Hameister H, Knöchel W, et al: SHARP is a novel component of the Notch/RBP-Jkappa signalling pathway. EMBO J. 21:5417–5426. 2002. View Article : Google Scholar : PubMed/NCBI

125 

Ariyoshi M and Schwabe JW: A conserved structural motif reveals the essential transcriptional repression function of Spen proteins and their role in developmental signaling. Genes Dev. 17:1909–1920. 2003. View Article : Google Scholar : PubMed/NCBI

126 

Legare S, Cavallone L, Mamo A, Chabot C, Sirois I, Magliocco A, Klimowicz A, Tonin PN, Buchanan M, Keilty D, et al: The estrogen receptor cofactor SPEN functions as a tumor suppressor and candidate biomarker of drug responsiveness in hormone-dependent breast cancers. Cancer Res. 75:4351–4363. 2015. View Article : Google Scholar : PubMed/NCBI

127 

Legare S, Chabot C and Basik M: SPEN, a new player in primary cilia formation and cell migration in breast cancer. Breast Cancer Res. 19:1042017. View Article : Google Scholar : PubMed/NCBI

128 

Margueron R and Reinberg D: The Polycomb complex PRC2 and its mark in life. Nature. 469:343–349. 2011. View Article : Google Scholar : PubMed/NCBI

129 

Wang X, Brea LT and Yu J: Immune modulatory functions of EZH2 in the tumor microenvironment: Implications in cancer immunotherapy. Am J Clin Exp Urol. 7:85–91. 2019.PubMed/NCBI

130 

Zingg D, Debbache J, Peña-Hernández R, Antunes AT, Schaefer SM, Cheng PF, Zimmerli D, Haeusel J, Calçada RR, Tuncer E, et al: EZH2-mediated primary cilium deconstruction drives metastatic melanoma formation. Cancer Cell. 34:69–84.e14. 2018. View Article : Google Scholar : PubMed/NCBI

131 

Denhez F, Wilcox-Adelman SA, Baciu PC, Saoncella S, Lee S, French B, Neveu W and Goetinck PF: Syndesmos, a syndecan-4 cytoplasmic domain interactor, binds to the focal adhesion adaptor proteins paxillin and Hic-5. J Biol Chem. 277:12270–12274. 2002. View Article : Google Scholar : PubMed/NCBI

132 

Avolio R, Jarvelin AI, Mohammed S, Agliarulo I, Condelli V, Zoppoli P, Calice G, Sarnataro D, Bechara E, Tartaglia GG, et al: Protein syndesmos is a novel RNA-binding protein that regulates primary cilia formation. Nucleic Acids Res. 46:12067–12086. 2018.PubMed/NCBI

133 

Haigis KM: KRAS alleles: The devil is in the detail. Trends Cancer. 3:686–697. 2017. View Article : Google Scholar : PubMed/NCBI

134 

Kempf E, Rousseau B, Besse B and Paz-Ares L: KRAS oncogene in lung cancer: Focus on molecularly driven clinical trials. Eur Respir Rev. 25:71–76. 2016. View Article : Google Scholar : PubMed/NCBI

135 

Pupo E, Avanzato D, Middonti E, Bussolino F and Lanzetti L: KRAS-driven metabolic rewiring reveals novel actionable targets in cancer. Front Oncol. 9:8482019. View Article : Google Scholar : PubMed/NCBI

136 

Eser S, Schnieke A, Schneider G and Saur D: Oncogenic KRAS signalling in pancreatic cancer. Br J Cancer. 111:817–822. 2014. View Article : Google Scholar : PubMed/NCBI

137 

Raleigh DR, Choksi PK, Krup AL, Mayer W, Santos N and Reiter JF: Hedgehog signaling drives medulloblastoma growth via CDK6. J Clin Invest. 128:120–124. 2018. View Article : Google Scholar : PubMed/NCBI

138 

Farooqi AA, de la Roche M, Djamgoz MBA and Siddik ZH: Overview of the oncogenic signaling pathways in colorectal cancer: Mechanistic insights. Semin Cancer Biol. 58:65–79. 2019. View Article : Google Scholar : PubMed/NCBI

139 

Palicelli A, Croci S, Bisagni A, Zanetti E, De Biase D, Melli B, Sanguedolce F, Ragazzi M, Zanelli M, Chaux A, et al: What do we have to know about PD-L1 expression in prostate cancer? a systematic literature review. Part 3: PD-L1, intracellular signaling pathways and tumor microenvironment. Int J Mol Sci. 22:123302021. View Article : Google Scholar : PubMed/NCBI

140 

Eguether T, Cordelieres FP and Pazour GJ: Intraflagellar transport is deeply integrated in hedgehog signaling. Mol Biol Cell. 29:1178–1189. 2018. View Article : Google Scholar : PubMed/NCBI

141 

Deng YZ, Cai Z, Shi S, Jiang H, Shang YR, Ma N, Wang JJ, Guan DX, Chen TW, Rong YF, et al: Cilia loss sensitizes cells to transformation by activating the mevalonate pathway. J Exp Med. 215:177–195. 2018. View Article : Google Scholar : PubMed/NCBI

142 

Jenks AD, Vyse S, Wong JP, Kostaras E, Keller D, Burgoyne T, Shoemark A, Tsalikis A, de la Roche M, Michaelis M, et al: Primary cilia mediate diverse kinase inhibitor resistance mechanisms in cancer. Cell Rep. 23:3042–3055. 2018. View Article : Google Scholar : PubMed/NCBI

143 

Khan NA, Willemarck N, Talebi A, Marchand A, Binda MM, Dehairs J, Rueda-Rincon N, Daniels VW, Bagadi M, Thimiri Govinda Raj DB, et al: Identification of drugs that restore primary cilium expression in cancer cells. Oncotarget. 7:9975–9992. 2016. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

August-2022
Volume 61 Issue 2

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Yin F, Wei Z, Chen F, Xin C and Chen Q: Molecular targets of primary cilia defects in cancer (Review). Int J Oncol 61: 98, 2022
APA
Yin, F., Wei, Z., Chen, F., Xin, C., & Chen, Q. (2022). Molecular targets of primary cilia defects in cancer (Review). International Journal of Oncology, 61, 98. https://doi.org/10.3892/ijo.2022.5388
MLA
Yin, F., Wei, Z., Chen, F., Xin, C., Chen, Q."Molecular targets of primary cilia defects in cancer (Review)". International Journal of Oncology 61.2 (2022): 98.
Chicago
Yin, F., Wei, Z., Chen, F., Xin, C., Chen, Q."Molecular targets of primary cilia defects in cancer (Review)". International Journal of Oncology 61, no. 2 (2022): 98. https://doi.org/10.3892/ijo.2022.5388