Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 1748 word(s) 1748 2021-08-23 10:40:39 |
2 update layout and reference Meta information modification 1748 2021-09-02 04:55:16 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Evason, K. β-Catenin in Hepatocellular Carcinoma. Encyclopedia. Available online: https://encyclopedia.pub/entry/13809 (accessed on 28 March 2024).
Evason K. β-Catenin in Hepatocellular Carcinoma. Encyclopedia. Available at: https://encyclopedia.pub/entry/13809. Accessed March 28, 2024.
Evason, Kimberley. "β-Catenin in Hepatocellular Carcinoma" Encyclopedia, https://encyclopedia.pub/entry/13809 (accessed March 28, 2024).
Evason, K. (2021, September 01). β-Catenin in Hepatocellular Carcinoma. In Encyclopedia. https://encyclopedia.pub/entry/13809
Evason, Kimberley. "β-Catenin in Hepatocellular Carcinoma." Encyclopedia. Web. 01 September, 2021.
β-Catenin in Hepatocellular Carcinoma
Edit

Hepatocellular carcinoma (HCC) is one of the deadliest human cancers. Activating mutations in the telomerase reverse transcriptase (TERT) promoter (TERTp) and CTNNB1 gene encoding β-catenin are widespread in HCC (~50% and ~30%, respectively). TERTp mutations are predicted to increase TERT transcription and telomerase activity. 

hepatocellular carcinoma TERT TERT promoter β-catenin

1. Introduction

Hepatocellular carcinoma (HCC) forms 80% of all primary liver cancers and is the fourth leading cause of cancer-related mortality globally. Effective treatments for this disease are limited due to its molecularly heterogeneous nature [1]. Telomerase reverse transcriptase (TERT), TP53 (tumor protein p53), and CTNNB1 encoding β-catenin are the top three most frequently mutated genes in HCC, altered in 47.1%, 29.2%, and 27.4% of cases from large-scale sequencing projects [2].
Both TERT and β-catenin play essential roles in liver development, tumorigenesis, and stemness pathways. In embryonic stem and progenitor cells, the TERT and WNT/β-catenin signaling axes crosstalk such that TERT enhances the transcription of β-catenin target genes, while β-catenin enhances the transcription of TERT by binding to its promoter region [3][4]. In HCC clinical specimens, activating β-catenin and TERT promoter (TERTp) mutations show significant concordance such that 74% of CTNNB1-mutated HCC also have TERTp mutations [5].
This CTNNB1-TERT association is special because CTNNB1 mutations are mutually exclusive to TP53 mutations [6] and to other perturbations in WNT pathway genes, such as AXIN1 and CCND1 [7]. The chromatin remodelers ARID2 (AT rich interactive domain 2) and ARID1A (AT rich interactive domain 1A) are associated with mutations in CTNNB1 (p-value = 0.005) and AXIN1 (p-value = 0.002), respectively [6]. Potential mechanisms underlying the ARID2-CTNNB1 and ARID1A-AXIN1 associations in liver cancer have not been explored in depth, perhaps because these genes are less commonly altered in HCC (~10%) than TERT or CTNNB1 [2].

2. Telomeres, Telomerase, and TERT

Telomeres are ribonucleoprotein caps composed of short repetitive non-coding DNA sequences and associated proteins that protect the ends of eukaryotic chromosomes [8]. Telomere length varies among species and serves to mitigate two biological complications. First, together with the associated Shelterin complex, telomeres help protect DNA near chromosome ends from the DNA damage response (DDR), which is induced by double-stranded breaks in the DNA and can lead to cell cycle arrest, apoptosis, or repair. Second, telomeres provide a cushion of non-coding DNA to postpone the end replication crisis, in which genome instability results from gradual shortening of the lagging strand in semi-conservative DNA replication. Due to the presence of telomeres, chromosomes can withstand 50–90 replication cycles while losing around 50–150 nucleotides per mitosis, even in the absence of telomere maintenance mechanisms [9].
Telomerase, a multi-subunit reverse transcriptase enzyme, plays a significant role in telomere maintenance. Telomerase levels are undetectable in most somatic cells; however, this enzyme is active in certain proliferating cells, germ cells, and the majority of cancers (~90%) [10]. The telomerase holoenzyme is composed of a catalytic protein subunit encoded by TERT, an RNA template component (TERC), and auxiliary proteins. TERT consists of four main domains: the TERT N-terminal domain (TEN), the TERT RNA binding domain (TRBD), a reverse transcriptase domain (RT), and a C-terminal extension domain. The TEN domain binds to the ssDNA of telomere overhangs at the 3′ end, the TRBD domain interacts with TERC, and the RT and C-terminal extension domains are involved in catalyzing the addition of telomere repeats to the 3′ end and binding to the resultant DNA/RNA hybrid [8][10]. In somatic cells, TERT is transcriptionally repressed via epigenetic mechanisms, mainly in the TERTp region [9].
TERT is also involved in many roles independent of its telomere lengthening activity [11]. These non-canonical functions of TERT include regulating gene expression, signal transduction, mitochondrial metabolism, and resistance to ionizing radiation [12]. Some of the most well-studied pathways regulated by TERT include the WNT/β-catenin pathway and the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) pathway, along with MYC, vascular endothelial growth factor (VEGF), and others [11]. TERT is found not only in the nucleus but also in the cytoplasm and mitochondria and is involved in tumorigenesis and cancer therapy resistance independent of telomere lengthening [12]. For example, the overexpression of mitochondrial TERT by HCC cells increases chemotherapeutic resistance (in vitro and in vivo) by increasing the mitochondrial membrane potential and inhibiting the mitochondrial apoptotic pathway [13]. The diverse roles of TERT are highlighted in Figure 1.
Figure 1. Telomerase reverse transcriptase (TERT) regulation and functions in cancer cells. Cancer cells may upregulate TERT by somatic TERTp mutations, TERT amplification and rearrangement, and/or epigenetic modifications in TERT and TERTp [8]. TERT performs both canonical (telomere lengthening) and non-canonical (gene regulation, signal transduction, stress response, and mitochondrial metabolism) functions [11][12]. Abbreviations: NF-κB, nuclear factor kappa-light-chain-enhancer of activated B cells; and VEGF, vascular endothelial growth factor.

3. Telomerase and Cancer

Without a telomere maintenance mechanism, cells undergoing oncogenesis and rapid divisions gradually suffer from telomere attrition, which eventually triggers replicative senescence or apoptosis [14]. To overcome this replication barrier caused by critical telomere length, cancer cells usually increase the TERT levels and/or TERT activity via (1) TERTp mutations, which increase TERT mRNA levels by creating new ETS binding motifs [15]; (2) amplification of TERT and/or TERC; (3) TERT rearrangement; and/or (4) epigenetic regulation of TERT and its promoter [8]. As cells in high-turnover tissues have higher telomerase activity, they may be able to stably upregulate TERT without acquiring activating mutations like TERTp mutations [14]. Supporting this possibility, cancers originating from tissues with high turnover rates tend to have a lower frequency of TERTp mutations than those from low-turnover tissues [14].
Thus, telomerase promotes telomere maintenance, and telomere maintenance is a cancer hallmark that allows malignant cells to overcome replicative senescence. TERT also promotes cellular transformation and proliferation independent of its telomerase activity [11]. Therefore, one would predict that loss-of-function mutations in TERT or TERC would inhibit tumorigenesis, while overexpression of TERT would promote tumor formation. Table 1 summarizes recent findings from cancer models with manipulation of Tert or Terc in mice or zebrafish. Supporting the aforementioned prediction, Tert overexpression in mice leads to an increased incidence of T-cell lymphomas [16], skin papillomas [17], and mammary tumors [18], depending on the promoter used to drive its expression (Table 1). Tert loss of function mutants display a delayed onset of mammary tumors [19], lymphomas [20], and early HCCs [21] (Table 1).
On the other hand, Terc knockdown in mice results in the converse effect: enhancement of tumor formation (Table 1). Terc−/− mice treated with carbon tetrachloride (CCl4) or diethylnitrosamine (DEN) display increased HCC “initiation foci”—microscopic lesions with atypical cytologic or architectural features that suggest malignancy but are not sufficiently well-developed to diagnose as HCC—although HCC number and size were decreased [22]. Similarly, loss-of-function mutation of zebrafish tert or terc resulted in an earlier onset of spontaneous tumors, including intestinal adenocarcinomas, hematopoietic neoplasms, hepatocellular adenomas, and germ cell tumors [23].
Why does telomerase seem to promote tumorigenesis in some animal models yet inhibit it in others? One possible explanation is that long-term inhibition of Tert or Terc, as in these genetic animal models, results in compensation by other telomere maintenance programs, such as alternative lengthening of telomeres (ALT). ALT is a telomerase-independent mechanism for telomere maintenance. ALT-mediated telomere synthesis occurs via homologous recombination between non-allelic telomeres, sister chromatids at telomeric sites, and/or linear or circular extrachromosomal telomeric repeats [24]. This compensatory mechanism might not be relevant to human hepatocarcinogenesis, during which TERT could potentially be inhibited more acutely using drugs. Inducible and/or conditional animal models in which Tert and/or Terc can be turned on at different time points during tumorigenesis would be helpful to dissect the importance of acute versus chronic telomerase inhibition.
Another possible reason why telomerase might inhibit tumorigenesis in some animal models while promoting it in others is related to the balance between canonical and non-canonical effects of Tert. Laboratory mouse models have long telomeres (25–40 kb vs. 10–15 kb in humans) that are likely sufficient to permit malignant transformation even in the absence of telomerase enzyme [12]. Therefore, Tert overexpression and mutant phenotypes in mouse tumor models may be at least partially independent of Tert telomere lengthening activity [16][17][18]. As discussed in the section above, non-canonical functions of TERT tend to be tumorigenic and anti-apoptotic. Thus, Tert deficiency in mice may delay tumorigenesis due to the resultant reduction in its non-canonical activity [19][21].
Zebrafish have telomere lengths similar to humans (~5–15 kb) [25] and, like humans, exhibit drastic shortening of telomeres with age [23]; therefore, the manipulation of tert levels is more likely to impact telomere length than in mice. Thus, the canonical effects of tert may be relatively more important to its role in zebrafish carcinogenesis. Zebrafish tert deficiency promotes the initiation of tumorigenesis through shortened telomeres that induce DNA damage responses, inflammation, and genetic instability [23]. A similar phenomenon occurs in aged late-generation Terc−/− mice, wherein shortened telomeres are associated with ulcerated skin lesions, defective wound healing, and genomic instability [26].
These findings highlight that TERT influences tumorigenesis in both canonical and non-canonical ways. Whether the net effects of TERT are pro- or anti-tumor depends on the stage of disease, tissue type, telomere biology in different animal models, when during tumor formation TERT is manipulated, and other factors. Additional studies of vertebrate HCC models, including rigorous examination of telomere length during carcinogenesis, could help define the importance of TERT’s telomerase activity to HCC pathogenesis.
Table 1. TERT cancer models in mice and zebrafish.
Gene (Animal) Site Specificity Expression * Result ** Ref.
Tert (mouse) Thymocytes and peripheral T cells + ↑ T-Cell Lymphomas [16]
Tert (mouse) Basal keratinocytes + ↑ skin papillomas (DMBA + TPA induction) [17]
Tert (mouse) Whole body + ↑ mammary tumors in aged females [18]
tert and terc (zebrafish) Neural progenitor cells + ↓ aggressiveness of RAS-mediated brain tumors [27]
Tert (mouse) Whole body Delayed onset of lymphomas in EμMYC mice [20]
Tert (mouse) Whole body Delayed onset of mammary tumors in PyMT mice [19]
Tert (mouse) Whole body ↓ HCC “initiation foci” (CCl4 induction) [21]
Terc (mouse) Whole body ↑ HCC “initiation foci” but ↓ HCC progression (uPA, CCl4 or DEN induction) [22]
Terc (mouse) Whole body ↑ tumors (lymphomas, teratocarcinomas, HCC, squamous cell carcinoma) [26]
Terc (mouse) Whole body ↓ skin papillomas (DMBA + TPA induction) [28]
Terc (mouse) Whole body ↑ epithelial cancers in TP53−/− mice [29]
Terc (mouse) Whole body ↑ adenoma initiation but ↓ progression in ApcMin mice [30]
terthu3430 (zebrafish) Whole body Earlier onset of tumors (germ cell tumors, hematopoietic neoplasms, HCA, etc.) [23]
terthu3430 (zebrafish) Whole body ↑ tumor incidence and aggressiveness of melanoma model *** [31]

*: +, gene overexpression; −, gene knockout or knockdown. **: ↑, increased; ↓, decreased. *** mitfa: HRAS (gives rise to melanomas) blastula cells transplanted into tert−/− casper embryos. Abbreviations: DMBA, 7,12-dimethylbenz[a]anthracene; TPA, 12-o-tetradecanoylphorbol 13-acetate; PyMT, polyomavirus middle T oncogene; CCl4, carbon tetrachloride; uPA, urokinase plasminogen activator; DEN, diethylnitrosamine; HCA, hepatocellular adenoma; Apc, adenomatous polyposis coli; ApcMin, multiple intestinal neoplasia (mutant) allele of Apc gene; and terthu3430, tert mutant line (allele hu3430) with a non-sense mutation resulting in a premature stop codon in exon 2 of tert gene.

References

  1. Yang, J.D.; Hainaut, P.; Gores, G.J.; Amadou, A.; Plymoth, A.; Roberts, L.R. A global view of hepatocellular carcinoma: Trends, risk, prevention and management. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 589–604.
  2. Lee, J.-S. The mutational landscape of hepatocellular carcinoma. Clin. Mol. Hepatol. 2015, 21, 220–229.
  3. Park, J.-I.; Venteicher, A.S.; Hong, J.Y.; Choi, J.; Jun, S.; Shkreli, M.; Chang, W.; Meng, Z.; Cheung, P.; Ji, H.; et al. Telomerase modulates Wnt signalling by association with target gene chromatin. Nature 2009, 460, 66–72.
  4. Hoffmeyer, K.; Raggioli, A.; Rudloff, S.; Anton, R.; Hierholzer, A.; Del Valle, I.; Hein, K.; Vogt, R.; Kemler, R. Wnt/β-Catenin Signaling Regulates Telomerase in Stem Cells and Cancer Cells. Science 2012, 336, 1549.
  5. Nault, J.C.; Mallet, M.; Pilati, C.; Calderaro, J.; Bioulac-Sage, P.; Laurent, C.; Laurent, A.; Cherqui, D.; Balabaud, C.; Zucman-Rossi, J. High frequency of telomerase reverse-transcriptase promoter somatic mutations in hepatocellular carcinoma and preneoplastic lesions. Nat. Commun. 2013, 4, 2218.
  6. Schulze, K.; Imbeaud, S.; Letouzé, E.; Alexandrov, L.B.; Calderaro, J.; Rebouissou, S.; Couchy, G.; Meiller, C.; Shinde, J.; Soysouvanh, F.; et al. Exome sequencing of hepatocellular carcinomas identifies new mutational signatures and potential therapeutic targets. Nat. Genet. 2015, 47, 505–511.
  7. Totoki, Y.; Tatsuno, K.; Covington, K.R.; Ueda, H.; Creighton, C.J.; Kato, M.; Tsuji, S.; Donehower, L.A.; Slagle, B.L.; Nakamura, H.; et al. Trans-ancestry mutational landscape of hepatocellular carcinoma genomes. Nat. Genet. 2014, 46, 1267–1273.
  8. Gaspar, T.B.; Sá, A.; Lopes, J.M.; Sobrinho-Simões, M.; Soares, P.; Vinagre, J. Telomere Maintenance Mechanisms in Cancer. Genes 2018, 9, 241.
  9. Heidenreich, B.; Kumar, R. TERT promoter mutations in telomere biology. Mutat. Res. Rev. Mutat. Res. 2017, 771, 15–31.
  10. Zhang, Q.; Kim, N.-K.; Feigon, J. Architecture of human telomerase RNA. Proc. Natl. Acad. Sci. USA 2011, 108, 20325–20332.
  11. Ségal-Bendirdjian, E.; Geli, V. Non-canonical Roles of Telomerase: Unraveling the Imbroglio. Front. Cell Dev. Biol. 2019, 7, 332.
  12. Romaniuk, A.; Paszel-Jaworska, A.; Totoń, E.; Lisiak, N.; Hołysz, H.; Królak, A.; Grodecka-Gazdecka, S.; Rubiś, B. The non-canonical functions of telomerase: To turn off or not to turn off. Mol. Biol. Rep. 2019, 46, 1401–1411.
  13. Yan, J.; Zhou, Y.; Chen, D.; Li, L.; Yang, X.; You, Y.; Ling, X. Impact of mitochondrial telomerase over-expression on drug resistance of hepatocellular carcinoma. Am. J. Transl. Res. 2015, 7, 88–99.
  14. Jafri, M.A.; Ansari, S.A.; Alqahtani, M.H.; Shay, J.W. Roles of telomeres and telomerase in cancer, and advances in telomerase-targeted therapies. Genome Med. 2016, 8, 69.
  15. Yuan, X.; Dai, M.; Xu, D. TERT promoter mutations and GABP transcription factors in carcinogenesis: More foes than friends. Cancer Lett. 2020, 493, 1–9.
  16. Canela, A.; Martín-Caballero, J.; Flores, J.M.; Blasco, M.A. Constitutive expression of tert in thymocytes leads to increased incidence and dissemination of T-cell lymphoma in Lck-Tert mice. Mol. Cell Biol. 2004, 24, 4275–4293.
  17. González-Suárez, E.; Samper, E.; Ramírez, A.; Flores, J.M.; Martín-Caballero, J.; Jorcano, J.L.; Blasco, M.A. Increased epidermal tumors and increased skin wound healing in transgenic mice overexpressing the catalytic subunit of telomerase, mTERT, in basal keratinocytes. EMBO J. 2001, 20, 2619–2630.
  18. Artandi, S.E.; Alson, S.; Tietze, M.K.; Sharpless, N.E.; Ye, S.; Greenberg, R.A.; Castrillon, D.H.; Horner, J.W.; Weiler, S.R.; Carrasco, R.D.; et al. Constitutive telomerase expression promotes mammary carcinomas in aging mice. Proc. Natl. Acad. Sci. USA 2002, 99, 8191–8196.
  19. Khattar, E.; Kumar, P.; Liu, C.Y.; Akıncılar, S.C.; Raju, A.; Lakshmanan, M.; Maury, J.J.P.; Qiang, Y.; Li, S.; Tan, E.Y.; et al. Telomerase reverse transcriptase promotes cancer cell proliferation by augmenting tRNA expression. J. Clin. Investig. 2016, 126, 4045–4060.
  20. Koh, C.M.; Khattar, E.; Leow, S.C.; Liu, C.Y.; Muller, J.; Ang, W.X.; Li, Y.; Franzoso, G.; Li, S.; Guccione, E.; et al. Telomerase regulates MYC-driven oncogenesis independent of its reverse transcriptase activity. J. Clin. Investig. 2015, 125, 2109–2122.
  21. Farazi, P.A.; Glickman, J.; Horner, J.; DePinho, R.A. Cooperative Interactions of p53 Mutation, Telomere Dysfunction, and Chronic Liver Damage in Hepatocellular Carcinoma Progression. Cancer Res. 2006, 66, 4766.
  22. Farazi, P.A.; Glickman, J.; Jiang, S.; Yu, A.; Rudolph, K.L.; DePinho, R.A. Differential Impact of Telomere Dysfunction on Initiation and Progression of Hepatocellular Carcinoma. Cancer Res. 2003, 63, 5021.
  23. Carneiro, M.C.; Henriques, C.M.; Nabais, J.; Ferreira, T.; Carvalho, T.; Ferreira, M.G. Short Telomeres in Key Tissues Initiate Local and Systemic Aging in Zebrafish. PLoS Genet. 2016, 12, e1005798.
  24. Marzec, P.; Armenise, C.; Pérot, G.; Roumelioti, F.-M.; Basyuk, E.; Gagos, S.; Chibon, F.; Déjardin, J. Nuclear-receptor-mediated telomere insertion leads to genome instability in ALT cancers. Cell 2015, 160, 913–927.
  25. Cayuela, M.L.; Claes, K.B.M.; Ferreira, M.G.; Henriques, C.M.; van Eeden, F.; Varga, M.; Vierstraete, J.; Mione, M.C. The Zebrafish as an Emerging Model to Study DNA Damage in Aging, Cancer and Other Diseases. Front. Cell Dev. Biol. 2019, 6, 178.
  26. Rudolph, K.L.; Chang, S.; Lee, H.-W.; Blasco, M.; Gottlieb, G.J.; Greider, C.; DePinho, R.A. Longevity, Stress Response, and Cancer in Aging Telomerase-Deficient Mice. Cell 1999, 96, 701–712.
  27. Idilli, A.I.; Cusanelli, E.; Pagani, F.; Berardinelli, F.; Bernabé, M.; Cayuela, M.L.; Poliani, P.L.; Mione, M.C. Expression of tert Prevents ALT in Zebrafish Brain Tumors. Front. Cell Dev. Biol. 2020, 8, 65.
  28. González-Suárez, E.; Samper, E.; Flores, J.M.; Blasco, M.A. Telomerase-deficient mice with short telomeres are resistant to skin tumorigenesis. Nat. Genet. 2000, 26, 114–117.
  29. Artandi, S.E.; Chang, S.; Lee, S.-L.; Alson, S.; Gottlieb, G.J.; Chin, L.; DePinho, R.A. Telomere dysfunction promotes non-reciprocal translocations and epithelial cancers in mice. Nature 2000, 406, 641–645.
  30. Rudolph, K.L.; Millard, M.; Bosenberg, M.W.; DePinho, R.A. Telomere dysfunction and evolution of intestinal carcinoma in mice and humans. Nat. Genet. 2001, 28, 155–159.
  31. Lex, K.; Maia Gil, M.; Lopes-Bastos, B.; Figueira, M.; Marzullo, M.; Giannetti, K.; Carvalho, T.; Ferreira, M.G. Telomere shortening produces an inflammatory environment that increases tumor incidence in zebrafish. Proc. Natl. Acad. Sci. USA 2020, 117, 15066.
More
Information
Subjects: Oncology
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 375
Revisions: 2 times (View History)
Update Date: 02 Sep 2021
1000/1000