Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Nat Rev Cancer. Author manuscript; available in PMC 2013 Oct 5.
Published in final edited form as:
PMCID: PMC3791171
NIHMSID: NIHMS514651
PMID: 23429735

Molecular pathogenesis and mechanisms of thyroid cancer

Abstract

Thyroid cancer is a common endocrine malignancy. There has been exciting progress in understanding its molecular pathogenesis in recent years, as best exemplified by the elucidation of the fundamental role of several major signalling pathways and related molecular derangements. Central to these mechanisms are the genetic and epigenetic alterations in these pathways, such as mutation, gene copy-number gain and aberrant gene methylation. Many of these molecular alterations represent novel diagnostic and prognostic molecular markers and therapeutic targets for thyroid cancer, which provide unprecedented opportunities for further research and clinical development of novel treatment strategies for this cancer.

Thyroid cancer is a common endocrine malignancy that has rapidly increased in global incidence in recent decades1,2. In the United States, the average annual increase in thyroid cancer incidence of 6.6% between 2000 and 2009 is the highest among all cancers2. Although the death rate of thyroid cancer is relatively low, the rate of disease recurrence or persistence is high, which is associated with increased incurability and patient morbidity and mortality3.

There are several histological types and subtypes of thyroid cancer with different cellular origins, characteristics and prognoses4 (TABLE 1). There are two types of endocrine thyroid cells — follicular thyroid cells and parafollicular C cells — from which thyroid cancers are derived. Follicular thyroid cell-derived tumours, including papillary thyroid cancer (PTC), follicular thyroid cancer (FTC), poorly differentiated thyroid cancer (PDTC) and anaplastic thyroid cancer (ATC), account for the majority of thyroid malignancies. PTC and FTC are collectively classified as differentiated thyroid cancer (DTC). Parafollicular C cell-derived medullary thyroid cancer (MTC) accounts for a small proportion of thyroid malignancies2. The primary molecular mechanism underlying MTC tumorigenesis is the aberrant activation of RET signalling (which is caused by RET mutations5), which are not present in follicular thyroid cell-derived tumours. The molecular pathogenesis of follicular thyroid cell-derived tumours is the focus of this Review.

Table 1

Thyroid tumours and their characteristics

Tumour typeCell of originPrevalence (% of thyroid cancers)Standard care and prognosisCharacteristics
FTAFollicular thyroid cells (which produce thyroid hormone and thyroglobulin)This is a benign lesionConservative monitoring; thyroidectomy if symptomaticCommon benign thyroid tumour; similar architecture to FTC, but typically encapsulated; lacking capsular or vascular invasion; lacking metastasis; lacking nuclear features of PTC
PTC*Follicular thyroid cells80–85Thyroidectomy and, in selected cases, radioiodine ablation (novel drugs for resistant disease); good overall prognosisWell differentiated, with papillary architecture and characteristic nuclear features that include enlargement, oval shape, elongation, overlapping and clearing, inclusions and grooves; propensity for lymphatic metastasis; PTC subtypes include conventional PTC (CPTC), follicular-variant PTC (FVPTC), tall-cell PTC (TCPTC) and a few rare variants
FTC*Follicular thyroid cells10–15Thyroidectomy and radioiodine ablation (novel drugs for resistant cases); good overall prognosisWell differentiated, hypercellular, microfollicular patterns, lacking nuclear features of PTC; vascular or capsular invasion; propensity for metastasis via the blood stream; Hürthle cell thyroid cancer is a unique subtype of FTC that accounts for 2–3% of thyroid cancers and is characterized by large, mitochondria-rich oncocytic cells and dense nuclei and nucleoli, as well as a high propensity for metastasis and a poor prognosis
PDTCFollicular thyroid cells5–10Surgery, radioiodine (in selected cases), chemotherapy, radiotherapy, novel drugs; poor prognosisPoorly differentiated, often overlapping with PTC and FTC; intermediate aggressiveness between differentiated and undifferentiated thyroid cancers
ATCFollicular thyroid cells2–3Surgery, chemotherapy, radiotherapy, novel drugs, palliative care; highly and rapidly lethalUndifferentiated; admixture of spindle, pleomorphic giant and epithelioid cells; extremely invasive and metastatic; highly lethal; may occur de novo or derive from PTC, FTC or PDTC
Medullary thyroid cancerParafollicular C cells (which produce calcitonin)2–3Surgery, chemotherapy, radiotherapy, novel drugs (for example, vandetanib)Moderate aggressiveness, high propensity for lymphatic metastasis; RET mutation; occurring in familial, MEN2 or sporadic forms
Primary lymphoma of the thyroid glandLymphocytes<1ChemotherapyUnusual and uncommon type of lymphoma
Metastatic cancer from other organsNon-thyroid origin<1Thyroidectomy in selected cases; treatment of original cancerMost commonly metastasized from renal and breast cancers; characteristics of original cancer

ATC, anaplastic thyroid cancer; FTA, follicular thyroid adenoma; FTC, follicular thyroid cancer; MEN2, multiple endocrine neoplasia type 2; PDTC, poorly differentiated thyroid cancer; PTC, papillary thyroid cancer.

*PTC and FTC are collectively classified as differentiated thyroid cancer (DTC).

Conventional surgical thyroidectomy with adjuvant ablation by radioiodine treatment has been the mainstay of treatment for follicular thyroid cell-derived cancer, but it is often not curative. The recent progress in understanding the molecular pathogenesis of thyroid cancer has shown great promise for the development of more-effective treatment strategies for thyroid cancer. This has mainly resulted from the identification of molecular alterations, including the genetic and epigenetic alterations of signalling pathways — such as the RAS–RAF–MEK–MAPK–ERK pathway (MAPK pathway) and the PI3K–AKT pathway — which is reshaping thyroid cancer medicine. This Review discusses the recent remarkable progress in understanding the molecular pathogenesis and mechanisms of thyroid cancer.

Common genetic alterations in thyroid cancer

Gene mutations

Numerous genetic alterations that have a fundamental role in the tumorigenesis of various thyroid tumours have been identified (TABLE 2). A prominent example is the T1799A transverse point mutation of BRAF, which results in the expression of BRAF-V600E mutant protein and causes constitutive activation of this serine/threonine kinase611. BRAFV600E mutation occurs in approximately 45% of PTCs12. There are also a few rare types of BRAF mutations identified in PTC, which mostly affect nucleotides around codon 600 and constitutively activate the BRAF kinase13,14. The requirement for BRAF-V600E to maintain tumour growth was initially demonstrated in a xenograft tumour model15. A previous comprehensive multicentre study demonstrated a strong association of BRAFV600E with poor clinicopathological outcomes of PTC, including aggressive pathological features, increased recurrence, loss of radioiodine avidity and treatment failures16. This was subsequently confirmed in numerous other studies, although, perhaps owing to variations in study design, some studies showed inconsistent results, as summarized and discussed in large meta-analyses that uniformly support the aggressive role of BRAFV600E mutation in PTC17,18. BRAFV600E-transgenic mice also developed aggressive PTC1921. Interestingly, some human PTC tumours have recently been found to exhibit intra-tumour heterogeneity in the BRAF genotype — with a minority of cells harbouring BRAFV600E and the majority harbouring wild-type BRAF22. This raises an interesting ‘chicken and egg’ puzzle of whether BRAF-V600E initiates PTC tumorigenesis or instead BRAFV600E occurs following the development of a thyroid tumour23. Although it is possible that BRAFV600E could be a secondary genetic event in PTC tumorigenesis as previously proposed24, an alternative possibility is that once PTC is initiated by BRAFV600E, secondary oncogenic alterations could take over to drive the tumorigenesis of PTC such that BRAF mutation is no longer selected for23.

Table 2

Gene mutations in thyroid tumours

MutationsTypes of thyroid tumoursApproximate prevalence (%)*Primary signalling pathways affectedFunctional impact on the protein and tumourRefs
BRAFV600ECPTC45MAPKActivating; promoting tumorigenesis, invasion, metastasis, recurrence and mortality612, 1618
FVPTC15
TCPTC80–100
ATC25
BRAFK601EFVPTC5MAPKActivating; probably similar to BRAFV600E14
HRAS, KRAS, NRASFTA20–25MAPK and PI3K–AKTActivating; promoting tumorigenesis, invasion and metastasis of PDTC and FTC26,31,33, 35,148, 184187, 206211
FTC30–45
FVPTC30–45
PDTC20–40
ATC20–30
PTEN (mutation)FTA0PI3K–AKTInactivating the gene but activating the PI3K pathway; promoting tumorigenesis and invasiveness26,3133, 212
FTC10–15
ATC10–20
PTC1–2
PTEN (deletion)FTC30PI3K–AKTInactivating the gene but activating the PI3K pathway; promoting tumorigenesis and invasiveness212
PIK3CAFTA0–5PI3K–AKTActivating; promoting tumorigenesis and invasiveness25,26, 3135
FTC5–15
ATC15–25
PTC1–2
AKT1Metastatic cancer15PI3K–AKTUnclear; seems to favour metastasis35
CTNNB1PDTC25WNT–β-cateninActivating; promoting tumour progression37,38
ATC60–65
TP53PDTC25p53-coupled pathwaysInactivating; promoting tumour progression39,40
ATC70–80
IDH1FTC5–25IDH1-associated metabolic pathwaysInactivating; impact on tumours is unclear41,42
FVPTC20
CPTC10
ATC10–30
ALKATC10MAPK and PI3K–AKTActivating; probably promoting tumour progression43
EGFRCPTC5MAPK and PI3K–AKTActivating; impact on tumours is unclear44
NDUFA13 (also known as GRIM19)HCTC15Component of complex I of the mitochondrial respiratory chainPresumably inactivating; affecting mitochondrial metabolism and cell death45

ALK, anaplastic lymphoma kinase; ATC, anaplastic thyroid cancer; CPTC, conventional PTC; CTNNB1, β-catenin; EGFR, epidermal growth factor receptor; FTA, follicular thyroid adenoma; FTC, follicular thyroid cancer; FVPTC, follicular-variant PTC; HCTC, Hürthle cell thyroid cancer; IDH1, isocitrate dehydrogenase 1; NDUFA13, NADH dehydrogenase (ubiquinone) 1α subcomplex 13; PDTC, poorly differentiated thyroid cancer; PTC, papillary thyroid cancer; TCPTC, tall-cell PTC.

*The values represent estimated overall prevalence of the indicated mutations.

Second in prevalence to BRAF mutations in thyroid cancer are RAS mutations. RAS is in its active state when bound with GTP. The intrinsic GTPase of RAS hydrolyses GTP and converts RAS into an inactive GDP-bound state, thus terminating RAS signalling. RAS mutations cause the loss of its GTPase activity, thus locking RAS in a constitutively active GTP-bound state. There are three isoforms of RAS: HRAS, KRAS and NRAS, and NRAS is predominantly mutated in thyroid tumours, mostly involving codons 12 and 61. Although RAS is a classical dual activator of the MAPK and PI3K–AKT pathways, RAS mutations seem to preferentially activate the PI3K–AKT pathway in thyroid tumorigenesis, as suggested by the preferential association of RAS mutations with AKT phosphorylation in thyroid cancers25,26. The common occurrence of RAS mutations in follicular thyroid adenoma (FTA), a presumed premalignant lesion, suggests that activated RAS may have a role in early follicular thyroid cell tumorigenesis. However, additional genetic alterations other than RAS mutation are apparently required to transform FTA into thyroid cancer. This is consistent with studies in which the expression of mutant HRAS was induced in normal thyroid cells in vitro, which resulted in only well-demarcated and differentiated colonies with phenotypes consistent with FTA27 and cell proliferation with normal differentiation28. More direct evidence has come from transgenic mouse studies in which conditional physiological expression of a KRAS mutant in the thyroid gland could not transform thyroid cells, but concurrent KRAS mutant expression and Pten deletion induced a rapid occurrence of aggressive FTC29.

Mutations or deletions of the tumour suppressor gene PTEN are the classical genetic alterations that activate the PI3K–AKT pathway and are the genetic basis for follicular thyroid cell tumorigenesis in Cowden’s syndrome30. Mutations of PIK3CA, which encodes the p110α catalytic subunit of PI3K, are also common in thyroid cancer, particularly FTC, PDTC and ATC25,26,3135. As in other human cancers36, activating PIK3CA mutations in thyroid cancer occur in exon 9 and exon 20. AKT1 mutations were found in metastatic thyroid cancers in one study, and their functional relevance remains to be established35.

Other important genes that are mutated in thyroid tumorigenesis include β-catenin (CTNNB1)37,38, TP5339,40, isocitrate dehydrogenase 1 (IDH1)41,42, anaplastic lymphoma kinase (ALK)43 and epidermal growth factor receptor (EGFR)44. The preferential occurrence of these mutations in PDTC and ATC, which are the most aggressive thyroid cancers, suggests that they may have a role in the progression and aggressiveness of thyroid cancer. In Hürthle-cell thyroid carcinoma (HCTC), mutations of NADH dehydrogenase (ubiquinone) 1α subcomplex 13 (NDUFA13; also known as GRIM19) are fairly common45. Unlike other types of thyroid cancers, HCTC does not harbour classical genetic alterations, such as BRAF and RAS mutations or RET–PTC 46,47, but commonly harbours DNA haploidization in recurrent disease47.

Gene amplifications and copy-number gains

Oncogenic gene amplification or copy-number gain are additional prominent genetic mechanisms in thyroid tumori-genesis (TABLE 3). This is particularly the case for genes encoding receptor tyrosine kinases (RTKs)26. Copy-number gains are also common for the genes encoding PI3K–AKT pathway members, including PIK3CA, PIK3CB, 3-phosphoinositide-dependent protein kinase 1 (PDPK1; also known as PDK1), AKT1 and AKT2 (REFS 25,26,32). Overall, genetic copy-number gains of these genes are more prevalent in ATC than in DTC, suggesting that these genetic alterations are important for the progression and aggressiveness of thyroid cancer. Copy-number gain of these genes in ATC could occur either through genetic amplification or chromosomal instability and aneuploidy. An important and expected consequence of these copy-number gains is the activation of downstream signalling pathways, as is suggested by the association of copy-number gains of genes encoding RTKs with increased phosphorylation of AKT and ERK in thyroid cancer26.

Table 3

Copy-number gains in thyroid cancer*

GenePrevalence in ATC cases (%)Prevalence in FTC cases (%)
EGFR19/41 (46.3)19/59 (32.2)
PDGFRA11/46 (23.9)4/52 (7.7)
PDGFRB14/37 (37.8)8/59 (13.6)
VEGFR120/44 (45.5)26/59 (44.1)
VEGFR28/46 (17.4)2/52 (3.8)
KIT10/46 (21.7)6/61 (9.8)
MET5/42 (11.9)5/58 (8.6)
PIK3CA18/47 (38.3)15/63 (23.8)
PIK3CB16/42 (38.1)25/55 (45.5)
PDPK1 (also known as PDK1)8/40 (20)14/58 (24.1)
AKT19/48 (18.8)5/61 (8.2)
AKT20/44 (0)13/58 (22.4)

ATC, anaplastic thyroid cancer; EGFR, epidermal growth factor receptor; FTC, follicular thyroid cancer; PDGFR, platelet-derived growth factor receptor; PDPK1, 3-phosphoinositide dependent protein kinase 1; VEGFR, vascular endothelial growth factor receptor.

*From REF. 26. Copy-number gains of some of these genes in thyroid cancers were also investigated in the studies reported in REFS 25,31–33, which showed comparable prevalences.

Many of the genes with copy-number gains are proto-oncogenes. A mechanism for them to contribute to thyroid tumorigenesis is through increased protein expression and consequent aberrant activation of the signalling pathways in which they are involved25,26,33. It is interesting to note that in DTCs, PIK3CA mutation is mutually exclusive with PIK3CA copy-number gain31,33, suggesting that either of these genetic alterations is sufficient to promote thyroid tumorigenesis through the PI3K–AKT pathway. However in aggressive undifferentiated ATC, PIK3CA mutations and amplifications often occur simultaneously in the same tumour26,33, which is a mechanism through which mutant PIK3CA can be amplified to drive the progression and aggressiveness of thyroid cancer.

Gene amplification or copy-number gain of the IQ-motif-containing GTPase-activating protein 1 (IQGAP1) is another important genetic event that has been recently discovered in thyroid cancer48. IQGAP1 is a multifunctional scaffold protein that has an important role in human tumorigenesis49. IQGAP1 amplification was found to increase protein expression and to be associated with invasiveness of thyroid cancer cells. Interestingly, coexistence of IQGAP1 copy-number gain and BRAFV600E mutation was particularly associated with a high incidence of recurrent PTC48.

Gene translocations

Gene translocation resulting in oncogenic rearrangements in thyroid cancer is best exemplified by RET–PTC. There are more than 10 types of RET–PTC translocation, as determined by the types of partner genes, and the most common types are RET–PTC1 and RET–PTC35052. RET is a proto-oncogene encoding an RTK. RET–PTC occurs as a consequence of genetic recombination between the 3′ tyrosine kinase portion of RET and the 5′ portion of a partner gene, such as the coiled-coil domain-containing gene 6 (CCDC6; also known as H4) in RET–PTC1 and nuclear receptor co-activator 4 (NCOA4; also known as ELE1) in RET–PTC3 (REFS 53,54). Spatial contiguity of RET and the partner gene in the nucleus is a structural basis for the formation of RET–PTC rearrangements55. The rearrangements result in ligand-independent dimerization and constitutive tyrosine kinase activity of RET. RET–PTC can occur in benign FTA and follicular variant PTC (FVPTC), but more commonly in classical PTC56. A recent study found a correlation between the presence of RET–PTC and a high growth rate of benign thyroid tumours57. However, the role of RET–PTC in early thyroid tumorigenesis is unclear. RET–PTC is a classical oncoprotein that activates the MAPK and PI3K–AKT pathways58,59. RET–PTC activates both pathways by recruiting signalling adaptors to phosphorylated Tyr1062 on the intracellular domain of the RET fusion protein60. It is thus not surprising that RET–PTC occurs in FTA and FVPTC; these are follicular thyroid tumours in which the PI3K–AKT pathway has a primary role in tumorigenesis61.

The paired box 8 (PAX8)–peroxisome proliferator-activated receptor-γ (PPARG) fusion gene (PAX8–PPARG) is another prominent recombinant oncogene in thyroid cancer, occurring in up to 60% of FTC and FVPTC6264. PAX8–PPARG also occurs in FTA64, although, like RET–PTC, its prevalence is low in benign thyroid tumours and its oncogenic role is unclear. PAX8–PPARγ exerts a dominant-negative effect on the wild-type tumour suppressor PPARγ and also transactivates certain PAX8-responsive genes65. Interestingly, the expression of PPARγ was dramatically reduced in FTC that developed in the TRβPV mouse model — which expresses a knock-in dominant-negative mutant of human thyroid hormone receptor-β (TRβ) — resulting in follicular thyroid tumorigenesis66. An AKAP9–BRAF fusion gene resulting in activation of the BRAF kinase also occurs in ionizing-radiation-induced PTC67, but not in sporadic PTC68. Unlike RET–PTC, the importance of AKAP9–BRAF in thyroid tumorigenesis is probably limited owing to its rarity.

Aberrant gene methylation

Aberrant gene methylation is an epigenetic hallmark of human cancer — which is also common in thyroid cancer69 — that usually silences a gene when occurring in the promoter regions. BRAFV600E mutation was found to be associated with hypermethylation of several tumour suppressor genes, including tissue inhibitor of metalloproteinases 3 (TIMP3), SLC5A8, death-associated protein kinase 1 (DAPK1) and retinoic acid receptor-β (RARB)70. A recent DNA methylation microarray study revealed broad hypermethylation of genes throughout the genome driven by BRAF-V600E signalling in PTC cells71. Interestingly, this study also revealed a large range of genes throughout the genome that, driven by BRAF-V600E, became hypomethylated and hence overexpressed. These hypermethylated or hypomethylated genes have important established metabolic and cellular functions. Thus, alterations in gene methylation coupled to BRAF-V600E and probably to other oncoproteins represent a prominent epigenetic mechanism in thyroid tumorigenesis.

Promoter methylation of PTEN is also common in FTC and ATC7274, which is consistent with the loss of expression of PTEN that is found in a subset of these thyroid cancers73,75,76. PTEN methylation is associated with genetic alterations of the PI3K–AKT pathway in thyroid tumours, including mutations of various isoforms of RAS, PIK3CA mutation and amplification, and PTEN mutations72. This is consistent with a model in which aberrant activation of the PI3K–AKT pathway, driven by activating genetic alterations, causes aberrant methylation and hence silencing of PTEN, which in turn leads to failure to terminate the PI3K–AKT signalling and creates a self-amplifying loop72.

Despite the identification of these common genetic alterations in thyroid cancers, it is important to note that approximately 30–35% of DTCs do not harbour known genetic alterations, and so further investigation is required to identify the underlying genetic backgrounds.

Altered signalling pathways in thyroid cancer

The MAPK signalling pathway

The MAPK pathway has a fundamental role in the regulation of cell proliferation and survival and in human tumorigenesis (FIG. 1). The importance of this pathway has been well established in thyroid tumorigenesis, particularly for PTC77,78. In thyroid cancer, the MAPK pathway is driven by activating mutations, including BRAF and RAS mutations, by RET–PTC and, in some cases, by the recently discovered ALK mutations43.

An external file that holds a picture, illustration, etc.
Object name is nihms514651f1.jpg
The MAPK and related pathways in thyroid cancer

Shown in the middle of the figure is the classical MAPK pathway leading from an extracellular mitogenic stimulus that activates a receptor tyrosine kinase (RTK) in the cell membrane, to RAS, RAF (shown as BRAF-V600E), MEK and ERK. ERK is activated by phosphorylation (P) and enters the nucleus where it upregulates tumour-promoting genes and downregulates tumour suppressor genes and thyroid iodide-handling genes. On the left side of the figure is the nuclear factor-κB (NF-κB) pathway, in which extracellular stimuli activate the pathway by acting on receptors in the cell membrane, leading to activation of the inhibitor of κB (IκB) kinase (IKK), resulting in the phosphorylation of IκB. Phosphorylated IκB becomes dissociated from NF-κB, which is normally bound with IκB in a complex and sequestered in the cytoplasm. Phosphorylated IκB undergoes ubiquitylation and proteasomal degradation. Free NF-κB then enters the nucleus to promote the expression of tumour-promoting genes. Through an undefined mechanism that is independent of MEK signalling, BRAF-V600E promotes the phosphorylation of IκB and the release of NF-κB, thus activating the NF-κB pathway. Shown on the right side of the figure is the RASSF1–mammalian STE20-like protein kinase 1 (MST1)–forkhead box O3 (FOXO3) pathway. Activated by extracellular pro-apoptotic stimuli through membrane receptors, RASSF1A activates MST1. Activated MST1 then phosphorylates FOXO3 on Ser207. The resulting phosphorylated FOXO3 becomes dissociated from 14-3-3 proteins in the cytoplasm. 14-3-3 proteins undergo proteasomal degradation, and phosphorylated FOXO3 enters the nucleus to promote the expression of pro-apoptotic genes in the FOXO pathway. BRAF-V600E directly interacts with and inhibits MST1 and prevents its activation by RASSF1A, thereby suppressing the pro-apoptotic signalling of the FOXO3 pathway. The downward arrow for the FOXO activities shown in the nucleus indicates this negative effect of BRAF-V600E on pro-apoptotic genes, which are normally upregulated by the RASSF1A–MST1–FOXO3 pathway. The triple independent coupling of BRAF-V600E to the pathways shown here represents a unique and powerful mechanism of thyroid tumorigenesis driven by BRAF-V600E. DAPK1, death-associated protein kinase 1; HIF1A, hypoxia-inducible factor 1α; MMP, matrix metalloproteinase; NIS, sodium–iodide symporter; TGFB1, transforming growth factor β1; TIMP3, tissue inhibitor of metalloproteinases 3; TPO, thyroid peroxidase; TSHR, thyroid-stimulating hormone receptor; TSP1, thrombospondin 1; UPA, urokinase plasminogen activator; UPAR, urokinase plasminogen activator receptor; VEGFA, vascular endothelial growth factor A.

MAPK-mediated thyroid tumorigenesis involves a wide range of secondary molecular alterations that synergize and amplify the oncogenic activity of this pathway, such as genome-wide hypermethylation and hypomethylation71. Additionally, upregulation of various well-established oncogenic proteins can occur. These include chemokines79,80, vascular endothelial growth factor A (VEGFA)81, MET82,83, nuclear factor-κB (NF-κB)84, matrix metalloproteinases (MMPs)79,84,85, prohibitin86, vimentin87, hypoxia-inducible factor 1α (HIF1α)88, prokineticin 1 (PROK1; also known as EG-VEGF)89, urokinase plasminogen activator (uPA) and its receptor (uPAR)90,91, transforming growth factor-β1 (TGFβ1)9294 and thrombospondin 1 (TSP1)93. These proteins are all established players in human tumorigenesis through a variety of mechanisms that drive cancer cell proliferation, growth, migration and survival, as well as tumour angiogenesis, invasion and metastasis.

Many of these proteins are key constituents of the unique extracellular matrix (ECM) microenvironments that have been proposed to be important in the pathogenesis of thyroid cancer driven by oncoproteins such as BRAF-V600E95. It is now clear that the ECM micro-environment does not only function as a structural support for the cellular elements of cancer, but it also has a profound impact on the behaviours of cancer cells, including their viability, proliferation, adhesion and motility. Thyroid cancer cells and stromal cells, such as fibroblasts and macrophages, produce proteins that form paracrine and autocrine signalling loops. For example, BRAF-V600E-mediated MAPK pathway activation promotes the release of TSP1 into the ECM where it interacts with and modulates other proteins. These include integrins and non-integrin cell-membrane receptors, matrix proteins, cytokines, VEGFA and MMPs, which in turn activate signalling in thyroid cancer cells and promote tumour progression and metastasis93,95. Another example is the tumour-promoting cytokine TGFβ1, which is released into the ECM owing to the activation of the BRAF-V600E–MAPK pathway92,94. Cytokines can create an inflammatory microenvironment that may produce reactive oxygen species and oxidative stress, which may in turn stimulate the MAPK pathway and augment thyroid tumorigenesis96. Thus, secondary molecular alterations in the tumour ECM microenvironment, triggered by the aberrant signalling of the MAPK pathway, have an important role in the pathogenesis of thyroid cancer.

In terms of stromal cells in the tumour micro-environment, an important role of tumour-associated macrophages is that they produce cytokines such as TGFβ1. The role of stromal fibroblasts in thyroid tumori-genesis seems to be affected by the expression patterns of the two fibroblast growth factor receptor 2 (FGFR2) isoforms, FGFR2-IIIb and FGFR2-IIIc, in cancer cells and fibroblasts: co-implantation of thyroid cancer cells and fibroblasts expressing the same type of FGFR2 isoform did not affect xenograft tumour progression, whereas co-implantation of thyroid cancer cells and fibroblasts that respectively expressed FGFR2-IIIb and FGFR2-IIIc resulted in increased tumour progression97. These results highlight the importance and complexity of cellular and molecular constituents in determining the impact of the microenvironment on thyroid cancer cell behaviour.

The PI3K–AKT signalling pathway

The PI3K–AKT pathway also has a fundamental role in thyroid tumori-genesis61,98 (FIG. 2). Early evidence to support the role of this pathway in thyroid tumorigenesis came from the finding that Cowden’s syndrome (which is caused by germline mutations of PTEN) was associated with FTA and FTC99. A functional role for the PI3K–AKT pathway in sporadic thyroid tumorigenesis was initially revealed by the finding of increased expression and activation of AKT in thyroid cancers, particularly FTC100102. Among the three AKT isoforms, AKT1 and AKT2, but not AKT3, were found to be robustly expressed and activated in thyroid cancer, suggesting a particularly important role for these two isoforms in thyroid tumorigenesis101.

An external file that holds a picture, illustration, etc.
Object name is nihms514651f2.jpg
The PI3K–AKT and related pathways in thyroid cancer

Extracellular signals activate receptor tyrosine kinases (RTKs) in the cell membrane, leading to the activation of RAS and subsequent activation of PI3K. Activated PI3K catalyses the conversion of phosphatidylinositol (4,5)-bisphosphate (PIP2) to phosphatidylinositol (3,4,5)-trisphosphate (PIP3). PIP3 activates 3-phosphoinositide-dependent protein kinase 1 (PDK1; also known as PDPK1), which consequently associates with AKT and leads to phosphorylation (P) and activation of AKT by PDK1. Phosphorylated AKT, which is an activated form of AKT, enters the nucleus where it induces tumour-promoting genes. In the cytoplasm, phospho-AKT also activates other signalling molecules or pathways, a prominent one being the mTOR pathway, which has an important role in tumorigenesis by promoting translation. Phospho-AKT can also directly phosphorylate glycogen synthase kinase 3β (GSK3β) and consequently inactivates it. GSK3β normally inhibits β-catenin, thus the effect of phospho-AKT is to relieve GSK3β-mediated suppression of β-catenin. Consequently, β-catenin can enter the nucleus where it promotes the expression of tumour-promoting genes. In the nucleus, phospho-AKT can phosphorylate forkhead box O3 (FOXO3) on its AKT-specific motif. Such phosphorylated FOXO3 is translocated out of the nucleus to the cytoplasm where it binds 14-3-3 proteins to be sequestered in the cytoplasm, thus terminating the pro-apoptotic activities of the FOXO3 pathway. The downward arrow for the FOXO activities in the nucleus in the figure indicates this negative effect of AKT on pro-apoptotic genes in the FOXO pathway, which would otherwise be upregulated by the FOXO3 pathway. This unique coupling of phospho-AKT to the three pathways provides a powerful driving force for thyroid tumorigenesis. The major negative regulatory mechanism of the PI3K–AKT pathway is PTEN, which is a phosphatase that converts PIP3 to PIP2, thus terminating the activation of the pathway. The inset shows the self enhancement mechanism of PI3K–AKT signalling in which genetic-alteration-driven activation of the pathway causes PTEN methylation and silencing with consequent loss of termination of the signalling, thus maintaining the pathway in full and constitutive activation.

In the TRβPV transgenic mouse model — in which the PI3K–AKT pathway is activated, which involves the interaction of TRβPV with the p85α regulatory subunit of PI3K — FTC spontaneously developed103,104, and the activation and nuclear localization of AKT1 was observed105. Human tumour studies suggested that the invasiveness and metastasis of FTC promoted by the PI3K–AKT pathway particularly involved the activation and nuclear localization of AKT1 (REF. 102). This is consistent with the occurrence of AKT1 mutations in metastatic thyroid cancers35. Furthermore, Akt1 ablation delayed or prevented tumour progression, vascular intravasation and distant metastasis of FTC in TRβPV mice106. The role of AKT2 in thyroid tumorigenesis has also been demonstrated in vivo, as the occurrence of thyroid tumours was reduced in Pten+/−Akt2−/− mice107. Thus, whereas the MAPK pathway has a central role primarily in PTC, the PI3K–AKT pathway has such a role primarily in FTC and its invasion and metastasis. This is consistent with genetic alterations in the PI3K–AKT pathway occurring most frequently in FTC, whereas those in the MAPK pathway occur most frequently in PTC17,18,61,78,98. The inhibition of thyroid cancer cell proliferation by PI3K–AKT pathway inhibitors was dependent on the presence of genetic alterations in the PI3K–AKT pathway108,109, suggesting that thyroid cancer cells harbouring such genetic alterations have become dependent on the over-activation of this pathway.

Vascular invasiveness and capsular invasiveness are defining histopathological features of FTC110. Genetic alterations in the PI3K–AKT pathway are far more common in FTC than in its precursor, FTA31,33. It is therefore conceivable that overactivated PI3K–AKT signalling promotes the conversion of FTA to FTC by conferring tumour cell invasiveness. There are also secondary molecular alterations, albeit limited in number so far, that are robustly induced by PI3K–AKT signalling and that have an important role in the pathogenesis of FTC; these include alterations of the WNT–β-catenin111, HIF1α112,113, FOXO3 (REF. 114) and NF-κB114 pathways.

The NF-κB signalling pathway

The NF-κB pathway has an important role in the regulation of inflammatory responses that are linked to tumorigenesis115 (FIG. 1). Increased NF-κB activation in thyroid cancer cell lines and tissues has long been documented116118. Recent studies demonstrate that the NF-κB pathway controls proliferative and anti-apoptotic signalling pathways in thyroid cancer cells119,120. The NF-κB pathway is involved in upregulating the expression of several oncogenic proteins that are also upregulated by the MAPK pathway. Furthermore, RET–PTC, RAS and BRAF-V600E — members of the MAPK pathway — can cause activation of the NF-κB pathway in thyroid cancers119. Interestingly, NF-κB signalling was upregulated by expression of BRAF-V600E121,122, but not by expression of a constitutively active mutant of MEK in NIH3T3 cells123, suggesting a direct coupling of BRAF-V600E to the NF-κB pathway. Through a mechanism that is not yet specifically defined, BRAF-V600E was shown to cause IκB degradation (and consequent activation of NF-κB) independently of MEK–ERK signalling in thyroid cancer cells84. This dual coupling of BRAF-V600E to the NF-κB pathway and to the MEK–ERK pathway is consistent with the finding that simultaneously suppressing the two pathways using NF-κB and MEK inhibitors synergistically inhibited the proliferation of thyroid cancer cells harbouring a BRAF-V600E mutation124.

The RASSF1–MST1–FOXO3 signalling pathway

RASSF1 is a member of the RAS association domain family, which, in response to apoptotic signalling, associates with and activates mammalian STE20-like protein kinase 1 (MST1; also known as STK4)125. Activated MST1 phosphorylates Ser207 in the forkhead domain of the forkhead transcription factor FOXO3. This phosphorylation disrupts the interaction of FOXO3 with 14-3-3 proteins in the cytoplasm and promotes FOXO3 translocation to the nucleus, where it promotes the transcription of pro-apoptotic genes126 (FIG. 1). Thus, this RASSF1–MST1–FOXO3 pathway has an important tumour-suppressive role by promoting apoptosis.

Hypermethylation of the promoter of RASSF1A is common and is associated with its silencing in thyroid cancers127,128. This occurs even in benign FTA, albeit to a lesser extent, suggesting that impairment of the RASSF1–MST1–FOXO3 pathway is involved in early thyroid tumorigenesis128.

Activation of the RASSF1–MST1–FOXO3 signalling pathway promoted thyroid cancer cell apoptosis21. Interestingly, this study demonstrated that BRAF-V600E, but not wild-type BRAF, directly interacted with the carboxyl terminus of MST1 and inhibited its kinase activities, resulting in decreased FOXO3 transactivation. This was independent of MEK–MAPK signalling. This suggests that, in addition to the classical coupling to MEK–MAPK, negative regulation of RASSF1–MST1–FOXO3 is a mechanism that is involved in BRAF-V600E-driven thyroid tumorigenesis. This unique mechanistic model explains the finding that thyroid cancers that developed in BRAFV600EMst1−/− transgenic mice were more aggressive than those that developed in BRAFV600E mice21. The direct interaction of BRAF-V600E with MST1 to prevent RASSF1-mediated MST1 activation also provides an explanation for the mutually exclusive occurrence of BRAFV600E mutation and RASSF1A hypermethylation in thyroid cancer128, as either event may be sufficient to inactivate RASSF1–MST1–FOXO3 signalling. Therefore, BRAF-V600E is coupled independently to three major pathways: MEK–MAPK, RASSF1–MST1–FOXO3 and NF-κB, and thus represents a unique and powerful oncogenic mechanism in thyroid tumorigenesis (FIG.1). The activation of the PI3K–AKT pathway, such as that induced by PTEN inactivation, can also downregulate the activities of the RASSF1–MST1–FOXO3 pathway in FTC114. This involves AKT-mediated phosphorylation of FOXO family members, which causes their translocation from the nucleus to the cytoplasm, where they are sequestered by 14-3-3 proteins, thus reducing the transcription of pro-apoptotic genes129 (FIG. 2).

The WNT–β-catenin signalling pathway

The WNT–β-catenin pathway has a well established role in the regulation of cell growth and proliferation, as well as in stem cell differentiation, and its constitutive activation is commonly found in human cancers130. β-catenin, when upregulated by upstream WNT signalling, is translocated into the nucleus and transcribes various tumour-promoting genes. In thyroid cancer, activation of WNT–β-catenin signalling is often caused by activating mutations of CTNNB1 (which encodes β-catenin), particularly in PDTC and ATC37,38. Furthermore, the expression of β-catenin was higher in ATC than in DTC131. Thus, the WNT–β-catenin pathway seems to have a particularly important role in thyroid tumour aggressiveness.

Aberrant activation of WNT–β-catenin signalling often occurs as a consequence of the activation of the PI3K–AKT pathway in thyroid cancers132,133. This occurs through glycogen synthase kinase 3β (GSK3β), which is directly phosphorylated — and hence inactivated — by AKT134. As GSK3β promotes the degradation of β-catenin, its inactivation results in the upregulation of WNT–β-catenin signalling (FIG. 2). Interestingly, RET–PTC can activate the WNT–β-catenin pathway by activating the PI3K–AKT pathway and also by directly phosphorylating β-catenin in thyroid cancer cells58.

The HIF1α pathway

It has long been known that hypoxia is a strong stimulus of cancer metabolism, growth and progression. HIF1α is a key mediator of the response to hypoxia, in which it binds to HIF1β (also known as ARNT) to form the HIF1 transcription factor that induces the expression of various genes associated with cell metabolism and tumour angiogenesis135. Angiogenesis, which is a key step in the progression of solid tumours, is a common response to intratumoural hypoxia. This process is promoted by VEGFA, the expression of which is strongly upregulated by HIF1. HIF1α is not expressed in normal thyroid tissues but is expressed in thyroid cancers, particularly in aggressive types, such as ATC, and this is consistent with a role in thyroid cancer progression112,136. The oncogene MET, which is another target of HIF1, is also abundantly expressed in association with upregulated HIF1α in thyroid cancer136. Interestingly, HIF1 can be upregulated in thyroid cancer by both the PI3K–AKT112,113 and the MAPK pathways88, thus contributing to the effects of these two pathways on thyroid tumour progression.

The thyroid-stimulating hormone receptor signalling pathway

Through activation by thyroid-stimulating hormone (TSH), the TSH receptor (TSHR) has a fundamental role in the regulation of thyroid cell proliferation, differentiation and function, as well as in the development of the thyroid gland137. TSHR is a guanine-nucleotide-binding G-protein-coupled receptor that triggers two intracellular signalling pathways: G-mediated adenylyl cyclase–cyclic AMP (cAMP) signalling and the Gq- or G11-mediated phospholipase Cβ–inositol 1,4,5-trisphosphate–intracellular Ca2+ signalling. FTC spontaneously developed in TRβPV mice, but when these mice were crossed with Tshr−/− mice, no thyroid cancer developed138. This study thus elegantly demonstrates the requirement of TSHR signalling in thyroid carcinogenesis in this mouse model. Similar findings were reported in another mouse model in which thyroid-specific knock-in of BrafV600E (LSL-BrafV600Ethyroid peroxidase (Tpo)-Cre) caused the development of aggressive PTC, and crossing these mice with Tshr−/− mice resulted in a failure to develop thyroid cancer139. In this study, thyroid-specific deletion of Gnas (which encodes G) in LSL-BrafV600ETpo-Cre mice attenuated thyroid tumour formation. Interestingly, there is a clinical association of higher serum levels of TSH with a higher risk for malignancy of thyroid nodules in humans140142.

However, it is not clear whether TSHR signalling is required directly for the initiation of thyroid cancer or whether it is simply required, as would be expected physiologically, for the TSHR-dependent generation of thyroid cells, from which an oncogene-driven thyroid cancer would originate. Overactivation of TSHR signalling, such as that achieved through activating mutations in TSHR or G, is well known to cause benign hyperfunctional FTA143,144. These tumours are almost never malignant, which suggests that TSHR signalling may be protective against malignant transformation of thyroid cells. Consistently, low serum levels of TSH are associated with common genetic variants that predispose to an increased risk of thyroid cancer145,146. It therefore seems that the TSH–TSHR system has a dichotomous role in the development of thyroid cancer: it may suppress malignant transformation of thyroid cells and hence suppress the occurrence of thyroid cancer, but it may promote the growth and progression of thyroid cancer once it has been initiated by oncogenic alterations.

Progressive molecular alterations

Progressive accumulation of genetic alterations

The accumulation of genetic alterations during thyroid tumour progression is probably best exemplified by genetic alterations to members of the PI3K–AKT pathway61. Mutations in members of this pathway, including those in RAS (particularly NRAS), PIK3CA and PTEN, as well as PIK3CA amplifications, increasingly occur from FTA to FTC and to ATC26,3133,147. Hypermethylation of the PTEN promoter also progressively increases in frequency from FTA to FTC and to ATC72. The coexistence of these genetic and epigenetic alterations also increasingly occurs from low-grade to high-grade thyroid tumours31,33. Genetic or epigenetic defects of PTEN can occur simultaneously with activating mutations of other genes in the PI3K–AKT pathway; in such cases, the PI3K–AKT pathway can presumably remain maximally activated26,33,72. Indeed, deletion of Pten accelerated the progression of FTC in TRβPV mice114.

The coexistence of multiple genetic alterations to members of the MAPK pathway also occurs, as exemplified by the simultaneous presence of BRAFV600E mutation, RAS mutations and RET–PTC in aggressive recurrent PTC and ATC26,148,149; these alterations are otherwise mutually exclusive in well-differentiated thyroid cancers7,9. It is thus compelling to propose that the process of thyroid cancer progression is one of progressive accumulation of multiple genetic alterations, which synergistically cooperate to amplify their oncogenicity.

Cooperation of the MAPK and PI3K–AKT pathways

The MAPK and PI3K–AKT pathways are primarily involved in differentiated PTC and FTC, respectively, and the simultaneous activation of both pathways becomes more frequent as the grade of thyroid tumours increases26,3133,35. One study analysed 24 genetic alterations in the major genes of the MAPK and PI3K–AKT pathways in 48 ATC samples and found that the majority (81%) of the samples harboured genetic alterations that could potentially activate both pathways26. The coexistence of phosphorylated ERK and AKT was also common in ATC but not in DTC26. Thus, genetic alterations that activate the MAPK and PI3K–AKT pathways are an important mechanism that drives the progression of thyroid cancer. In this model, as illustrated in FIG. 3, when a genetic alteration occurs in the MAPK pathway it drives tumorigenesis of the thyroid cell towards PTC; when a genetic alteration occurs in the PI3K–AKT pathway it drives tumorigenesis towards FTA and FTC. As genetic alterations accumulate and both pathways become activated, the tumour progresses into PDTC and ATC, which is a process that may be further accelerated by additional or secondary genetic alterations, including mutations in TP53, CTNNB1 and ALK.

An external file that holds a picture, illustration, etc.
Object name is nihms514651f3.jpg
Model of the progression of thyroid tumorigenesis driven by the MAPK and PI3K–AKT pathways

Activation of the MAPK pathway by genetic alterations, such as the BRAFV600E mutation, primarily drives the development of papillary thyroid cancer (PTC) from follicular thyroid cells. By contrast, activation of the PI3K–AKT pathway by genetic alterations, such as mutations in RAS, PTEN and PIK3CA, primarily drives the development of follicular thyroid adenoma (FTA) and follicular thyroid cancer (FTC) from follicular thyroid cells. Conversion from FTA to FTC is largely due to increasing activation of the PI3K–AKT pathway. As genetic alterations accumulate and intensify the signalling of each of the two pathways, PTC and FTC can progress to poorly differentiated thyroid cancer (PDTC). When both pathways are fully activated through accumulated genetic alterations, conversion from PDTC to anaplastic thyroid cancer (ATC) is strongly facilitated. It is also possible that PDTC and ATC can both develop de novo directly from follicular thyroid cells, and that ATC can develop from PTC or FTC if appropriate genetic alterations occur. Many secondary molecular alterations also progressively accumulate and synergize with the two pathways in driving the progression of thyroid tumorigenesis, as discussed in the text (not shown in the figure). The increasing number of vertical arrows and colour intensity of the ovals symbolize the increasing genetic alterations and signalling of the two pathways as thyroid tumorigenesis progresses. The figure is modified from REF.33 © (2007) American Association for Cancer Research.

This model is supported by studies of transgenic mice in which the deletion of Pten and knock-in of KrasG12D (to activate both MAPK and PI3K–AKT pathways), but not either genetic manipulation alone, caused the development of aggressive thyroid cancer29. Similarly, this phenomenon is observed in melanoma. Transgenic mice with induced expression of BRAFV600E in melanocytes to activate the MAPK pathway alone developed only melanocytic hyperplasias, but the expression of BRAFV600E with Pten deletion (to also activate the PI3K–AKT pathway) caused melanoma with a rapid onset of metastasis150. Thus, dual activation of the MAPK and PI3K–AKT pathways may be a common mechanism for promoting tumour progression.

Impairment of the iodide-handling machinery

The main and unique function of follicular thyroid cells is to use iodide to synthesize thyroid hormone151,152 (FIG. 4). In this process, iodide is transported into the cell through the sodium–iodide symporter (NIS) that is located in the basal membrane. At the apical membrane, pendrin transports iodide out of the cell into the lumen of the thyroid follicle. In the lumen of the thyroid follicle, iodide undergoes oxidation by TPO and is incorporated into tyrosine residues of thyroglobulin (TG), which is later cleaved through proteolysis to produce thyroid hormones. This process is upregulated by TSH-mediated activation of TSHR. This is the biological basis for the conventional radioiodine treatment of thyroid cancer, but the iodide-handling machinery is often impaired, particularly in advanced thyroid cancers, thus making radioiodine treatment ineffective.

An external file that holds a picture, illustration, etc.
Object name is nihms514651f4.jpg
Iodide-handling machinery in the thyroid cell and its silencing by BRAF-V600E

a | A follicular thyroid cell is shown, which abundantly expresses the molecules involved in the uptake and metabolism of iodide (I), including the sodium–iodide symporter (NIS) in the basal membrane, which transports I coupled with Na+ into the cell from the extracellular compartment. I is then transported into the follicular lumen via pendrin in the apical membrane where it is oxidized by thyroid peroxidase (TPO) and incorporated into tyrosine amino-acid residues in thyroglobulin (TG) to form iodinated TG (TG-I) for the synthesis of thyroid hormone. The whole process is upregulated by cyclic AMP (cAMP) signalling that is triggered by the binding of thyroid-stimulating hormone (TSH) to its receptor (TSHR) in the membrane. With normal expression and function of this system, I is abundantly taken up and accumulated in the follicular thyroid cell and in the follicular lumen. b | BRAF-V600E, through activating the MAPK pathway in thyroid cancer, causes the silencing of thyroid-specific genes and shuts off the iodide-handling machinery. Consequently, I uptake is reduced in the thyroid cell and is sparsely accumulated in the follicular lumen. For simplicity, much molecular detail is omitted.

Aberrant activation of the MAPK pathway has a crucial role in the impairment of the iodide-handling machinery17. BRAFV600E mutation is associated with the loss of radioiodine avidity and with radioiodine treatment failure in PTC16,35,153155. Consistently, BRAFV600E mutation is highly prevalent (78–95% of cases) in recurrent radioiodine-refractory PTC16,35,149,155, in contrast with the lower prevalence of BRAFV600E mutation (45% of cases) in primary PTC12. Numerous studies have reported an association of BRAFV600E mutation with decreased or absent expression of thyroid iodide-handling genes: NIS, TSHR, TPO, TG and SLC26A4 (which encodes pendrin) in thyroid cancer82,153,154,156160. A direct role of BRAFV600E in the downregulation of thyroid hormone synthesis pathway genes was demonstrated by showing that induced expression of BRAF-V600E in thyroid cells impaired the expression of almost all of these genes, which was restored by ceasing the expression of BRAF-V600E or by suppressing the MAPK pathway with a MEK inhibitor161. Similarly, in thyroid cells expressing RET–PTC1, NIS expression was increased following treatment with a MEK inhibitor162. Suppression of BRAF-V600E in mouse models of thyroid cancer also restored the expression of thyroid iodide-handling genes and radioiodine uptake163. This aberrant gene silencing by BRAF-V600E may involve alterations to histone acetylation at gene promoters164,165. An autocrine loop involving TGFβ is another mechanism92: BRAF-V600E promotes the secretion of TGFβ, which, through the activation of SMADs and consequent impairment of the thyroid-gene transcription factor PAX8, is a potent repressor of NIS in thyroid cells166.

Activation of the PI3K–AKT pathway was also shown to downregulate the iodide-handling machinery in thyroid cells both in vitro and in vivo167,168. Inhibition of the PI3K–AKT pathway induced NIS expression and iodide uptake in human thyroid cancer cells164,169. In addition to NIS, suppression of the PI3K–AKT pathway also induced the expression of TSHR, TPO and TG in human thyroid cancer cells, which was enhanced by treatment with histone deacetylase inhibitors, and radioiodine uptake was correspondingly robustly induced164. This could be augmented by co-treatment with TSH164. The involvement of both MAPK and PI3K–AKT pathways in the silencing of thyroid iodide-handling genes is consistent with the accumulation of genetic alterations in both pathways as thyroid tumours progress, during which there is an increasing loss of radioiodine avidity.

Translational promises in thyroid cancer

Remarkable advances in the translation of molecular findings in thyroid cancer to the clinic have occurred recently170. For example, because of its cancer specificity, the detection of BRAFV600E mutation in fine-needle aspiration biopsy (FNAB) samples was initially tested for the evaluation of thyroid nodules171,172, which has now found increasing clinical use170,173. The diagnostic utility of detecting RET–PTC using FNAB has also been intensively investigated170. The combinatorial use of genetic markers identified in FNAB samples, including BRAF mutation, RAS mutations, RET–PTC and PAX8–PPARG, has been shown to improve the diagnostic accuracy for thyroid nodules that are otherwise diagnostically indeterminate by conventional cytology assessment, although this diagnostic approach may need further improvement174,175. Applying genomic information to address clinical issues of thyroid neoplasia has also shown promise. For example, a gene-expression classifier (with high sensitivity and negative predictive power) is being used in the clinic to assist the diagnostic evaluation of indeterminate thyroid nodules176.

The prognostic application of BRAFV600E mutation has also become part of the clinical management of PTC170,173. A large, multicentre study demonstrated a strong association of BRAFV600E mutation with patient mortality from PTC177. Large studies have demonstrated that BRAFV600E is strongly associated with poor clinico-pathological outcomes even in conventionally low-risk patients16,178181. BRAFV600E detected in FNAB samples preoperatively predicted poor clinicopathological outcomes of PTC180,182,183 and even predicted lymph node metastasis (as determined by prophylactic central neck dissection in patients without preoperative evidence of lymph node metastasis)183. It is thus recommended that preoperative testing of BRAFV600E in FNAB samples be used for assisting risk stratification and defining surgical and medical treatments for patients with PTC170. Owing to their association with a worse prognosis of PDTC, RAS mutations are also promising prognostic markers for this type of thyroid cancer184186. It has been recently shown that RAS mutations are also associated with poor clinicopathological outcomes of FTC187 but this needs to be evaluated further.

Many components of the MAPK and PI3K–AKT pathways, from RTKs in the cell membrane to the various downstream signalling relay molecules, such as BRAF, MEK, AKT and mTOR, are therapeutic targets that are being actively tested for the treatment of thyroid cancer170. Novel small-molecule protein-kinase inhibitors have shown promise in clinical trials on thyroid cancer, including axitinib188, sorafenib189191, motesanib192 and pazopanib193. A promising therapeutic strategy is genetic-based targeting of thyroid cancer194, as supported by many preclinical studies demonstrating the selective inhibition of BRAFV600E-mutant thyroid cancer cells by various MEK inhibitors124,195199 and BRAF-V600E-specific inhibitors200,201. The latter include PLX4032 (also known as vemurafenib), which has been recently approved by the US Food and Drug Administration (FDA) for the treatment of BRAFV600E-positive melanoma202. Some of these MEK and BRAF-V600E inhibitors are currently being clinically tested for thyroid cancer. Therapeutic targeting of the PI3K–AKT pathway may also be genetically guided, as genetic alterations that activate this pathway confer thyroid cancer cells with remarkably increased sensitivities to AKT and mTOR inhibitors108,109. This therapeutic strategy for thyroid cancer may prove to be useful and should be tested clinically, as encouraged by a recent clinical trial showing that mutations in the PI3K–AKT pathway significantly increased the response rate of breast and gynaecological cancers to inhibitors of the PI3K–AKT pathway203.

The involvement of multiple signalling pathways in aggressive thyroid cancer suggests that it may be necessary to target them simultaneously for effective treatment. Indeed, single agent clinical trials of various kinase inhibitors in thyroid cancer have generally shown only partial responses. Several recent preclinical studies testing the combination of MEK or BRAF-V600E inhibitors with PI3K, AKT or mTOR inhibitors showed synergistic effects in inhibiting the proliferation and promoting apoptosis of thyroid cancer cells124,199,204,205. Synergistic effects of simultaneously targeting the MAPK and PI3K–AKT pathways were even more pronounced in cells that harboured genetic alterations in both pathways199. It has been recently demonstrated that simultaneously inhibiting the MAPK, PI3K–AKT and histone deacetylase pathways could more robustly induce the expression of thyroid iodide-handling genes (and thus increase radioiodine uptake) in thyroid cancer cells compared with inhibiting each pathway alone, and this could be enhanced by co-treatment with TSH164. Therefore, to restore radioiodine avidity by simultaneously targeting multiple signalling pathways is another promising area of translational research on thyroid cancer.

Concluding remarks

Remarkable progress in understanding the molecular pathogenesis of thyroid cancer has been made in recent years, as exemplified by the elucidation of the fundamental role of signalling pathways, such as the MAPK and PI3K–AKT pathways. Activation of these pathways, often in close connection and cooperation, constitute the primary oncogenic mechanism that promotes the development and progression of thyroid cancer. Central to this mechanism are the many powerful oncogenic alterations that drive these pathways. Important secondary molecular derangements driven by the overactivation of these pathways have also been increasingly uncovered, which synergize and amplify oncogenic signalling in thyroid tumorigenesis. This understanding of the molecular pathogenesis of thyroid cancer has now opened unprecedented opportunities for the development of novel clinical strategies for the management of thyroid cancer.

At a glance

  • Thyroid cancer is a common endocrine malignancy, and exciting progress has occurred in recent years in understanding its molecular pathogenesis.
  • Genetic and epigenetic alterations are the driving forces of thyroid cancer. Examples of these alterations include mutations in BRAF (BRAFV600E), RAS, PIK3CA, PTEN, TP53, β-catenin (CTNNB1), anaplastic lymphoma kinase (ALK) and isocitrate dehydrogenase 1 (IDH1), translocations (RET–PTC and paired box 8 (PAX8)–peroxisome proliferator-activated receptor-γ (PPARG)) and aberrant gene methylation.
  • At the core of the molecular pathogenesis of thyroid cancer is the uncontrolled activity of various signalling pathways, including the MAPK, PI3K–AKT, nuclear factor-κB (NF-κB), RASSF1–mammalian STE20-like protein kinase 1 (MST1)–forkhead box O3 (FOXO3), WNT–β-catenin, hypoxia-inducible factor 1α (HIF1α) and thyroid-stimulating hormone (TSH)–TSH receptor (TSHR) pathways.
  • The progression of thyroid cancer is a process of accumulation of genetic and epigenetic alterations with corresponding progressive derangements of signalling pathways. These are accompanied by numerous secondary molecular alterations, both in the cell and in the tumour microenvironment, which, acting in cooperation, amplify and synergize their impacts on thyroid tumorigenesis.
  • Aberrant silencing of thyroid iodide-handling genes and consequent loss of radioiodine avidity of thyroid cancer promoted by BRAF-V600E is a unique molecular pathological process in thyroid cancer, which causes the failure of radioiodine treatment.
  • The recent molecular findings provide unprecedented opportunities for further research and clinical development of novel molecular-based treatment strategies for thyroid cancer.

Acknowledgments

This work is supported by the US National Institutes of Health (NIH) R01 grants CA113507 and CA134225 to the author. The author thanks A. K. Murugan, Y. Trink and D. Liu for their help in preparing the figures and references. The author also thanks the numerous colleagues and investigators in this field whose outstanding work made it an extremely enjoyable experience to write this Review. The author wishes to apologize to those whose work is not cited owing to space limitations.

Glossary

Differentiated thyroid cancer(DTC). Collectively includes both papillary thyroid cancer (PTC) and follicular thyroid cancer (FTC), which are histologically differentiated, by comparison with poorly differentiated thyroid cancer (PDTC) and undifferentiated anaplastic thyroid cancer (ATC)
Radioiodine treatmentA classical treatment for patients with thyroid cancer. It is used after total thyroidectomy and consists of treating patients with the radioiodine isotope 131I, which emits high-energy β-particles and takes advantage of the unique function of thyroid follicular cells to accumulate iodide as a substrate for the synthesis of thyroid hormone
MAPK pathwayThis pathway has a fundamental role in the regulation of cell growth, proliferation, apoptosis and metabolic activities, through regulating the expression of various genes
PI3K–AKT pathwayThis pathway also has a fundamental role in the regulation of cell growth, proliferation, apoptosis and metabolic activities, through regulating the expression of various genes
PtenA tumour suppressor gene, the protein product of which converts phosphatidylinositol (3,4,5)-trisphosphate (PIP3) to phosphatidylinositol (4,5)-trisphosphate (PIP2), counteracting the conversion of PIP2 to PIP3 by PI3K and thus terminating PI3K–AKT signalling
Cowden’s syndromeAlso known as Cowden’s disease or multiple hamartoma syndrome. A rare, autosomally inherited disorder that is caused by mutations or defects in the tumour suppressor gene PTEN and is characterized by multiple tumour-like growths called hamartomas and an increased risk of certain cancers, such as breast cancer and follicular thyroid cancer
Copy-number gainA genetic abnormality in which the copy number of a chromosomal region or gene is more than the normal two copies (one paternal allele and one maternal allele), which occurs through the amplification of a local region of DNA within a chromosome, or through aneuploidy, in which multiple copies or fragments of identical chromosomes are present
Receptor tyrosine kinases(RTKs). These are typically growth-promoting transmembrane receptors that transduce extracellular signalling into intracellular signalling through the activation of their cytoplasmic kinase domain in response to extracellular signals such as ligand-binding
Gene translocationA genetic rearrangement in which a chromosomal fragment is translocated to another chromosome where it is not normally located, which may create a recombinant gene product with new function or uncontrolled function (compared with the original gene product)
Follicular thyroid tumoursCollectively includes follicular thyroid adenoma (FTA), follicular thyroid cancer (FTC) and follicular variant papillary thyroid cancer (FVPTC); they share common dense follicular cell architectures
Gene methylationAn epigenetic covalent addition of a methyl group to the fifth carbon of the cytosine residue in a CpG dinucleotide, typically in CpG islands (that is, CpG-rich regions) in the 5′-flanking promoter regions of genes. Such methylation is usually closely associated with chromatin remodelling and silencing of the corresponding genes. Capsular invasiveness, A phenomenon in which thyroid cancer invades the connective tissue capsule that surrounds the tumour, which is a defining feature of progression to malignancy from a benign thyroid tumour

Footnotes

Competing interests statement

The author declares competing financial interests: see Web version for details.

References

1. Jemal A, et al. Global cancer statistics. CA Cancer J Clin. 2011;61:69–90. [PubMed] [Google Scholar]
2. Howlader N, et al. SEER Cancer Statistics Review 1975–2009 (Vintage 2009 Populations) National Cancer Institute. 2012 [online], http://seer.cancer.gov/csr/1975_2009_pops09.
3. Tuttle RM, et al. Thyroid carcinoma. J Natl Compr Canc Netw. 2010;8:1228–1274. [PubMed] [Google Scholar]
4. DeLellis RA, Lloyd RV, Heitz PU, Eng C. World Health Organization Classification of Tumours. Pathology And Genetics Of Tumors Of Endocrine Organs. IARC Press; 2004. This book describes the most recent version of World Health Organization classification of thyroid tumours. [Google Scholar]
5. Hofstra RM, et al. A mutation in the RET proto-oncogene associated with multiple endocrine neoplasia type 2B and sporadic medullary thyroid carcinoma. Nature. 1994;367:375–376. [PubMed] [Google Scholar]
6. Cohen Y, et al. BRAF mutation in papillary thyroid carcinoma. J Natl Cancer Inst. 2003;95:625–627. This is the first submitted and accepted study that reported a high prevalence of BRAFV600E mutation in thyroid cancer, which was awarded a United States patent. [PubMed] [Google Scholar]
7. Soares P, et al. BRAF mutations and RET/PTC rearrangements are alternative events in the etiopathogenesis of PTC. Oncogene. 2003;22:4578–4580. [PubMed] [Google Scholar]
8. Namba H, et al. Clinical implication of hot spot BRAF mutation, V599E, in papillary thyroid cancers. J Clin Endocrinol Metab. 2003;88:4393–4397. [PubMed] [Google Scholar]
9. Kimura ET, et al. High prevalence of BRAF mutations in thyroid cancer: genetic evidence for constitutive activation of the RET/PTC-RAS-BRAF signaling pathway in papillary thyroid carcinoma. Cancer Res. 2003;63:1454–1457. References 7 and 9 demonstrate a mutual exclusivity between BRAF, RAS and RET–PTC mutations, thus suggesting an independent role of BRAF mutation in activating the MAPK pathway. [PubMed] [Google Scholar]
10. Xu X, Quiros RM, Gattuso P, Ain KB, Prinz RA. High prevalence of BRAF gene mutation in papillary thyroid carcinomas and thyroid tumor cell lines. Cancer Res. 2003;63:4561–4567. [PubMed] [Google Scholar]
11. Fukushima T, et al. BRAF mutations in papillary carcinomas of the thyroid. Oncogene. 2003;22:6455–6457. [PubMed] [Google Scholar]
12. Xing M. BRAF mutation in thyroid cancer. Endocr Relat Cancer. 2005;12:245–262. This is the first comprehensive review of BRAF mutation in thyroid cancer that provides a good summary and analysis of the prevalence and other characteristics of this mutation in various types of thyroid cancer. [PubMed] [Google Scholar]
13. Hou P, Liu D, Xing M. Functional characterization of the T1799-1801del and A1799-1816ins BRAF mutations in papillary thyroid cancer. Cell Cycle. 2007;6:377–379. [PubMed] [Google Scholar]
14. Trovisco V, et al. Type and prevalence of BRAF mutations are closely associated with papillary thyroid carcinoma histotype and patients’ age but not with tumour aggressiveness. Virchows Arch. 2005;446:589–595. [PubMed] [Google Scholar]
15. Liu D, Liu Z, Condouris S, Xing M. BRAF V600E maintains proliferation, transformation, and tumorigenicity of BRAF-mutant papillary thyroid cancer cells. J Clin Endocrinol Metab. 2007;92:2264–2271. [PMC free article] [PubMed] [Google Scholar]
16. Xing M, et al. BRAF mutation predicts a poorer clinical prognosis for papillary thyroid cancer. J Clin Endocrinol Metab. 2005;90:6373–6379. This is the first comprehensive multicentre study that shows the association of BRAF mutation with aggressive behaviours of PTC, including aggressive pathological features, disease recurrence, loss of radioiodine avidity of the tumour and treatment failure. [PubMed] [Google Scholar]
17. Xing M. BRAF mutation in papillary thyroid cancer: pathogenic role, molecular bases, and clinical implications. Endocr Rev. 2007;28:742–762. This is a large meta-analysis on the role of BRAF mutation in the aggressiveness of PTC. [PubMed] [Google Scholar]
18. Kim TH, et al. The association of the BRAF(V600E) mutation with prognostic factors and poor clinical outcome in papillary thyroid cancer: a meta-analysis. Cancer. 2012;118:1764–1773. This is a large meta-analysis on the relationship of BRAF mutation with clinicopathological behaviours of PTC. [PubMed] [Google Scholar]
19. Charles RP, Iezza G, Amendola E, Dankort D, McMahon M. Mutationally activated BRAF(V600E) elicits papillary thyroid cancer in the adult mouse. Cancer Res. 2011;71:3863–3871. [PMC free article] [PubMed] [Google Scholar]
20. Knauf JA, et al. Targeted expression of BRAFV600E in thyroid cells of transgenic mice results in papillary thyroid cancers that undergo dedifferentiation. Cancer Res. 2005;65:4238–4245. This is the first transgenic mouse study demonstrating that BRAF-V600E could induce the development of aggressive PTC. [PubMed] [Google Scholar]
21. Lee SJ, et al. Cross-regulation between oncogenic BRAF(V600E) kinase and the MST1 pathway in papillary thyroid carcinoma. PLoS ONE. 2011;6:e16180. This study for the first time demonstrated a direct role for BRAF-V600E in downregulating the pro-apoptotic RASSF1–MST1–FOXO3 signalling pathway by demonstrating the direct interaction of BRAF-V600E with MST1. [PMC free article] [PubMed] [Google Scholar]
22. Guerra A, et al. The primary occurrence of BRAF(V600E) is a rare clonal event in papillary thyroid carcinoma. J Clin Endocrinol Metab. 2012;97:517–524. [PubMed] [Google Scholar]
23. Xing M. BRAFV600E mutation and papillary thyroid cancer: chicken or egg? J Clin Endocrinol Metab. 2012;97:2295–2298. [PMC free article] [PubMed] [Google Scholar]
24. Vasko V, et al. High prevalence and possible de novo formation of BRAF mutation in metastasized papillary thyroid cancer in lymph nodes. J Clin Endocrinol Metab. 2005;90:5265–5269. [PubMed] [Google Scholar]
25. Abubaker J, et al. Clinicopathological analysis of papillary thyroid cancer with PIK3CA alterations in a Middle Eastern population. J Clin Endocrinol Metab. 2008;93:611–618. [PubMed] [Google Scholar]
26. Liu Z, et al. Highly prevalent genetic alterations in receptor tyrosine kinases and phosphatidylinositol 3-kinase/akt and mitogen-activated protein kinase pathways in anaplastic and follicular thyroid cancers. J Clin Endocrinol Metab. 2008;93:3106–3116. This study demonstrates a high prevalence of genetic alterations that can activate both the MAPK and PI3K–AKT pathways in aggressive thyroid cancer, thus providing a strong genetic basis for the joint role of the two pathways in the progression and aggressiveness of thyroid cancer. [PubMed] [Google Scholar]
27. Bond JA, Wyllie FS, Rowson J, Radulescu A, Wynford-Thomas D. In vitro reconstruction of tumour initiation in a human epithelium. Oncogene. 1994;9:281–290. [PubMed] [Google Scholar]
28. Gire V, Wynford-Thomas D. RAS oncogene activation induces proliferation in normal human thyroid epithelial cells without loss of differentiation. Oncogene. 2000;19:737–744. [PubMed] [Google Scholar]
29. Miller KA, et al. Oncogenic Kras requires simultaneous PI3K signaling to induce ERK activation and transform thyroid epithelial cells in vivo. Cancer Res. 2009;69:3689–3694. [PMC free article] [PubMed] [Google Scholar]
30. Gustafson S, Zbuk KM, Scacheri C, Eng C. Cowden syndrome. Semin Oncol. 2007;34:428–434. [PubMed] [Google Scholar]
31. Wang Y, et al. High prevalence and mutual exclusivity of genetic alterations in the phosphatidylinositol-3-kinase/akt pathway in thyroid tumors. J Clin Endocrinol Metab. 2007;92:2387–2390. [PubMed] [Google Scholar]
32. Santarpia L, El Naggar AK, Cote GJ, Myers JN, Sherman SI. Phosphatidylinositol 3-kinase/akt and ras/raf-mitogen-activated protein kinase pathway mutations in anaplastic thyroid cancer. J Clin Endocrinol Metab. 2008;93:278–284. [PubMed] [Google Scholar]
33. Hou P, et al. Genetic alterations and their relationship in the phosphatidylinositol 3-kinase/Akt pathway in thyroid cancer. Clin Cancer Res. 2007;13:1161–1170. This study demonstrates collectively common genetic alterations in the PI3K–AKT pathway in thyroid cancer, thus providing a strong genetic basis for the extensive role of this pathway in thyroid tumorigenesis. [PubMed] [Google Scholar]
34. Garcia-Rostan G, et al. Mutation of the PIK3CA gene in anaplastic thyroid cancer. Cancer Res. 2005;65:10199–10207. This is the first study that demonstrates common PIK3CA mutations in ATC. [PubMed] [Google Scholar]
35. Ricarte-Filho JC, et al. Mutational profile of advanced primary and metastatic radioactive iodine-refractory thyroid cancers reveals distinct pathogenetic roles for BRAF, PIK3CA, and AKT1. Cancer Res. 2009;69:4885–4893. [PMC free article] [PubMed] [Google Scholar]
36. Samuels Y, et al. High frequency of mutations of the PIK3CA gene in human cancers. Science. 2004;304:554. [PubMed] [Google Scholar]
37. Garcia-Rostan G, et al. Frequent mutation and nuclear localization of β-catenin in anaplastic thyroid carcinoma. Cancer Res. 1999;59:1811–1815. This is the first study to demonstrate common mutations in β-catenin in ATC. [PubMed] [Google Scholar]
38. Garcia-Rostan G, et al. β-catenin dysregulation in thyroid neoplasms: down-regulation, aberrant nuclear expression, and CTNNB1 exon 3 mutations are markers for aggressive tumor phenotypes and poor prognosis. Am J Pathol. 2001;158:987–996. [PMC free article] [PubMed] [Google Scholar]
39. Fagin JA, et al. High prevalence of mutations of the p53 gene in poorly differentiated human thyroid carcinomas. J Clin Invest. 1993;91:179–184. [PMC free article] [PubMed] [Google Scholar]
40. Donghi R, et al. Gene p53 mutations are restricted to poorly differentiated and undifferentiated carcinomas of the thyroid gland. J Clin Invest. 1993;91:1753–1760. References 39 and 40 demonstrate the high prevalence of mutations in TP53 in ATC. [PMC free article] [PubMed] [Google Scholar]
41. Murugan AK, Bojdani E, Xing M. Identification and functional characterization of isocitrate dehydrogenase 1 (IDH1) mutations in thyroid cancer. Biochem Biophys Res Commun. 2010;393:555–559. This is the first demonstration that IDH1 mutations occur in thyroid cancer. [PMC free article] [PubMed] [Google Scholar]
42. Hemerly JP, Bastos AU, Cerutti JM. Identification of several novel non-p. R132 IDH1 variants in thyroid carcinomas. Eur J Endocrinol. 2010;163:747–755. [PubMed] [Google Scholar]
43. Murugan AK, Xing M. Anaplastic thyroid cancers harbor novel oncogenic mutations of the ALK gene. Cancer Res. 2011;71:4403–4411. This is the first report of novel ALK mutations that can activate both the MAPK and PI3K–AKT pathways in ATC. [PMC free article] [PubMed] [Google Scholar]
44. Murugan AK, Dong J, Xie J, Xing M. Uncommon GNAQ, MMP8, AKT3, EGFR, and PIK3R1 mutations in thyroid cancers. Endocr Pathol. 2011;22:97–102. [PMC free article] [PubMed] [Google Scholar]
45. Maximo V, et al. Somatic and germline mutation in GRIM-19, a dual function gene involved in mitochondrial metabolism and cell death, is linked to mitochondrion-rich (Hurthle cell) tumours of the thyroid. Br J Cancer. 2005;92:1892–1898. [PMC free article] [PubMed] [Google Scholar]
46. Musholt PB, et al. Follicular histotypes of oncocytic thyroid carcinomas do not carry mutations of the BRAF hot-spot. World J Surg. 2008;32:722–728. [PubMed] [Google Scholar]
47. Corver WE, et al. Genome haploidisation with chromosome 7 retention in oncocytic follicular thyroid carcinoma. PLoS ONE. 2012;7:e38287. [PMC free article] [PubMed] [Google Scholar]
48. Liu Z, et al. IQGAP1 plays an important role in the invasiveness of thyroid cancer. Clin Cancer Res. 2010;16:6009–6018. [PMC free article] [PubMed] [Google Scholar]
49. White CD, Brown MD, Sacks DB. IQGAPs in cancer: a family of scaffold proteins underlying tumorigenesis. FEBS Lett. 2009;583:1817–1824. [PMC free article] [PubMed] [Google Scholar]
50. Rabes HM, et al. Pattern of radiation-induced RET and NTRK1 rearrangements in 191 post-chernobyl papillary thyroid carcinomas: biological, phenotypic, and clinical implications. Clin Cancer Res. 2000;6:1093–1103. [PubMed] [Google Scholar]
51. Santoro M, et al. Gene rearrangement and Chernobyl related thyroid cancers. Br J Cancer. 2000;82:315–322. [PMC free article] [PubMed] [Google Scholar]
52. Klugbauer S, Demidchik EP, Lengfelder E, Rabes HM. Molecular analysis of new subtypes of ELE/RET rearrangements, their reciprocal transcripts and breakpoints in papillary thyroid carcinomas of children after Chernobyl. Oncogene. 1998;16:671–675. [PubMed] [Google Scholar]
53. Ciampi R, Nikiforov YE. RET/PTC rearrangements & BRAF mutations in thyroid tumorigenesis. Endocrinology. 2007;148:936–941. [PubMed] [Google Scholar]
54. Santoro M, Melillo RM, Fusco A. RET/PTC activation in papillary thyroid carcinoma: European Journal of Endocrinology Prize Lecture. Eur J Endocrinol. 2006;155:645–653. [PubMed] [Google Scholar]
55. Nikiforova MN, et al. Proximity of chromosomal loci that participate in radiation-induced rearrangements in human cells. Science. 2000;290:138–141. This study demonstrates the proximity of chromosomal loci that are involved in the formation of RET–PTC translocations, thus providing a structural basis and explanation for the formation of genetic rearrangements. [PubMed] [Google Scholar]
56. Marotta V, Guerra A, Sapio MR, Vitale M. RET/ PTC rearrangement in benign and malignant thyroid diseases: a clinical standpoint. Eur J Endocrinol. 2011;165:499–507. [PubMed] [Google Scholar]
57. Sapio MR, et al. High growth rate of benign thyroid nodules bearing RET/PTC rearrangements. J Clin Endocrinol Metab. 2011;96:e916–e919. [PubMed] [Google Scholar]
58. Castellone MD, et al. The β-catenin axis integrates multiple signals downstream from RET/papillary thyroid carcinoma leading to cell proliferation. Cancer Res. 2009;69:1867–1876. [PMC free article] [PubMed] [Google Scholar]
59. Miyagi E, et al. Chronic expression of RET/PTC 3 enhances basal and insulin-stimulated PI3 kinase/AKT signaling and increases IRS-2 expression in FRTL-5 thyroid cells. Mol Carcinog. 2004;41:98–107. [PubMed] [Google Scholar]
60. Hayashi H, et al. Characterization of intracellular signals via tyrosine 1062 in RET activated by glial cell line-derived neurotrophic factor. Oncogene. 2000;19:4469–4475. [PubMed] [Google Scholar]
61. Xing M. Genetic alterations in the phosphatidylinositol-3 kinase/Akt pathway in thyroid cancer. Thyroid. 2010;20:697–706. This is a comprehensive review on genetic alterations in the PI3K–AKT pathway in thyroid cancer. [PMC free article] [PubMed] [Google Scholar]
62. Kroll TG, et al. PAX8-PPARγ1 fusion oncogene in human thyroid carcinoma. Science. 2000;289:1357–1360. [PubMed] [Google Scholar]
63. Dwight T, et al. Involvement of the PAX8/peroxisome proliferator-activated receptor γ rearrangement in follicular thyroid tumors. J Clin Endocrinol Metab. 2003;88:4440–4445. [PubMed] [Google Scholar]
64. Eberhardt NL, Grebe SK, McIver B, Reddi HV. The role of the PAX8/PPARγ fusion oncogene in the pathogenesis of follicular thyroid cancer. Mol Cell Endocrinol. 2010;321:50–56. [PMC free article] [PubMed] [Google Scholar]
65. Placzkowski KA, Reddi HV, Grebe SK, Eberhardt NL, McIver B. The role of the PAX8/ PPARγ fusion oncogene in thyroid cancer. PPAR Res. 2008;2008:672829. [PMC free article] [PubMed] [Google Scholar]
66. Ying H, et al. Mutant thyroid hormone receptor β represses the expression and transcriptional activity of peroxisome proliferator-activated receptor γ during thyroid carcinogenesis. Cancer Res. 2003;63:5274–5280. [PubMed] [Google Scholar]
67. Ciampi R, et al. Oncogenic AKAP9-BRAF fusion is a novel mechanism of MAPK pathway activation in thyroid cancer. J Clin Invest. 2005;115:94–101. [PMC free article] [PubMed] [Google Scholar]
68. Lee JH, Lee ES, Kim YS, Won NH, Chae YS. BRAF mutation and AKAP9 expression in sporadic papillary thyroid carcinomas. Pathology. 2006;38:201–204. [PubMed] [Google Scholar]
69. Xing M. Gene methylation in thyroid tumorigenesis. Endocrinology. 2007;148:948–953. [PubMed] [Google Scholar]
70. Hu S, et al. Association of aberrant methylation of tumor suppressor genes with tumor aggressiveness and BRAF mutation in papillary thyroid cancer. Int J Cancer. 2006;119:2322–2329. [PubMed] [Google Scholar]
71. Hou P, Liu D, Xing M. Genome-wide alterations in gene methylation by the BRAF V600E mutation in papillary thyroid cancer cells. Endocr Relat Cancer. 2011;18:687–697. This is a DNA methylation microarray study that revealed genome-wide hypomethylation and hypermethylation and hence expression alterations of genes driven by BRAFV600E, thus uncovering a prominent epigenetic mechanism in thyroid tumorigenesis driven by this oncogene. [PMC free article] [PubMed] [Google Scholar]
72. Hou P, Ji M, Xing M. Association of PTEN gene methylation with genetic alterations in the phosphatidylinositol 3-kinase/AKT signaling pathway in thyroid tumors. Cancer. 2008;113:2440–2447. This study proposes a model in which activation of the PI3K–AKT pathway, driven by genetic alterations, causes aberrant methylation and hence silencing of PTEN, which in turn leads to the failure in terminating PI3K–AKT signalling, thus creating a self-amplification loop that enhances and maintains the constitutive activation of the PI3K–AKT pathway in thyroid cancer. [PubMed] [Google Scholar]
73. Alvarez-Nunez F, et al. PTEN promoter methylation in sporadic thyroid carcinomas. Thyroid. 2006;16:17–23. [PubMed] [Google Scholar]
74. Schagdarsurengin U, Gimm O, Dralle H, Hoang-Vu C, Dammann R. CpG island methylation of tumor-related promoters occurs preferentially in undifferentiated carcinoma. Thyroid. 2006;16:633–642. [PubMed] [Google Scholar]
75. Frisk T, et al. Silencing of the PTEN tumor-suppressor gene in anaplastic thyroid cancer. Genes Chromosomes Cancer. 2002;35:74–80. [PubMed] [Google Scholar]
76. Bruni P, et al. PTEN expression is reduced in a subset of sporadic thyroid carcinomas: evidence that PTEN-growth suppressing activity in thyroid cancer cells mediated by p27kip1. Oncogene. 2000;19:3146–3155. [PubMed] [Google Scholar]
77. Kondo T, Ezzat S, Asa SL. Pathogenetic mechanisms in thyroid follicular-cell neoplasia. Nature Rev Cancer. 2006;6:292–306. [PubMed] [Google Scholar]
78. Xing M. Recent advances in molecular biology of thyroid cancer and their clinical implications. Otolaryngol Clin North Am. 2008;41:1135–1146. [PMC free article] [PubMed] [Google Scholar]
79. Melillo RM, et al. The RET/PTC-RAS-BRAF linear signaling cascade mediates the motile and mitogenic phenotype of thyroid cancer cells. J Clin Invest. 2005;115:1068–1081. [PMC free article] [PubMed] [Google Scholar] Retracted
80. Oler G, et al. Gene expression profiling of papillary thyroid carcinoma identifies transcripts correlated with BRAF mutational status and lymph node metastasis. Clin Cancer Res. 2008;14:4735–4742. [PMC free article] [PubMed] [Google Scholar]
81. Jo YS, et al. Influence of the BRAF V600E mutation on expression of vascular endothelial growth factor in papillary thyroid cancer. J Clin Endocrinol Metab. 2006;91:3667–3670. [PubMed] [Google Scholar]
82. Giordano TJ, et al. Molecular classification of papillary thyroid carcinoma: distinct BRAF, RAS, and RET/PTC mutation-specific gene expression profiles discovered by DNA microarray analysis. Oncogene. 2005;24:6646–6656. [PubMed] [Google Scholar]
83. Kumagai A, et al. Childhood thyroid carcinoma with BRAFT1799A mutation shows unique pathological features of poor differentiation. Oncol Rep. 2006;16:123–126. [PubMed] [Google Scholar]
84. Palona I, et al. BRAFV600E promotes invasiveness of thyroid cancer cells through nuclear factor κB activation. Endocrinology. 2006;147:5699–5707. This study demonstrates a direct coupling of BRAF-V600E to the NF-κB pathway that is independent of MEK–ERK signalling in thyroid cancer. [PubMed] [Google Scholar]
85. Mesa C, Jr, et al. Conditional activation of RET/PTC3 and BRAFV600E in thyroid cells is associated with gene expression profiles that predict a preferential role of BRAF in extracellular matrix remodeling. Cancer Res. 2006;66:6521–6529. [PubMed] [Google Scholar]
86. Franzoni A, et al. Prohibitin is overexpressed in papillary thyroid carcinomas bearing the BRAF(V600E) mutation. Thyroid. 2009;19:247–255. [PubMed] [Google Scholar]
87. Watanabe R, et al. Possible involvement of BRAFV600E in altered gene expression in papillary thyroid cancer. Endocr J. 2009;56:407–414. [PubMed] [Google Scholar]
88. Zerilli M, et al. BRAF(V600E) mutation influences hypoxia-inducible factor-1α expression levels in papillary thyroid cancer. Mod Pathol. 2010;23:1052–1060. [PubMed] [Google Scholar]
89. Pasquali D, et al. Upregulation of endocrine gland-derived vascular endothelial growth factor in papillary thyroid cancers displaying infiltrative patterns, lymph node metastases, and BRAF mutation. Thyroid. 2011;21:391–399. [PubMed] [Google Scholar]
90. Nowicki TS, et al. Inhibition of uPAR and uPA reduces invasion in papillary thyroid carcinoma cells. Laryngoscope. 2010;120:1383–1390. [PubMed] [Google Scholar]
91. Nowicki TS, et al. Downregulation of uPAR inhibits migration, invasion, proliferation, FAK/PI3K/Akt signaling and induces senescence in papillary thyroid carcinoma cells. Cell Cycle. 2011;10:100–107. [PMC free article] [PubMed] [Google Scholar]
92. Riesco-Eizaguirre G, et al. The BRAFV600E oncogene induces transforming growth factor β secretion leading to sodium iodide symporter repression and increased malignancy in thyroid cancer. Cancer Res. 2009;69:8317–8325. This study demonstrates a unique mechanism in which BRAF-V600E causes silencing of NIS by promoting the secretion of TGFβ, which in turn activates SMAD to silence NIS. [PubMed] [Google Scholar]
93. Nucera C, et al. B-Raf(V600E) and thrombospondin-1 promote thyroid cancer progression. Proc Natl Acad Sci USA. 2010;107:10649–10654. This study demonstrates an important role of TSP1 and its relationship with BRAF-V600E, thus providing an excellent example of the role of the tumour microenvironment in thyroid tumorigenesis. [PMC free article] [PubMed] [Google Scholar]
94. Knauf JA, et al. Progression of BRAF-induced thyroid cancer is associated with epithelial-mesenchymal transition requiring concomitant MAP kinase and TGFβ signaling. Oncogene. 2011;30:3153–3162. [PMC free article] [PubMed] [Google Scholar]
95. Nucera C, Lawler J, Parangi S. BRAF(V600E) and microenvironment in thyroid cancer: a functional link to drive cancer progression. Cancer Res. 2011;71:2417–2422. [PMC free article] [PubMed] [Google Scholar]
96. Xing M. Oxidative stress: a new risk factor for thyroid cancer. Endocr Relat Cancer. 2012;19:C7–C11. [PMC free article] [PubMed] [Google Scholar]
97. Guo M, Liu W, Serra S, Asa SL, Ezzat S. FGFR2 isoforms support epithelial-stromal interactions in thyroid cancer progression. Cancer Res. 2012;72:2017–2027. [PubMed] [Google Scholar]
98. Saji M, Ringel MD. The PI3K-Akt-mTOR pathway in initiation and progression of thyroid tumors. Mol Cell Endocrinol. 2010;321:20–28. [PMC free article] [PubMed] [Google Scholar]
99. Liaw D, et al. Germline mutations of the PTEN gene in Cowden disease, an inherited breast and thyroid cancer syndrome. Nature Genet. 1997;16:64–67. [PubMed] [Google Scholar]
100. Miyakawa M, et al. Increased expression of phosphorylated p70S6 kinase and Akt in papillary thyroid cancer tissues. Endocr J. 2003;50:77–83. [PubMed] [Google Scholar]
101. Ringel MD, et al. Overexpression and overactivation of Akt in thyroid carcinoma. Cancer Res. 2001;61:6105–6111. This study provides the first functional evidence for the involvement of the PI3K–AKT pathway in thyroid tumorigenesis and thyroid cancer invasion. [PubMed] [Google Scholar]
102. Vasko V, et al. Akt activation and localisation correlate with tumour invasion and oncogene expression in thyroid cancer. J Med Genet. 2004;41:161–170. [PMC free article] [PubMed] [Google Scholar]
103. Suzuki H, Willingham MC, Cheng SY. Mice with a mutation in the thyroid hormone receptor β gene spontaneously develop thyroid carcinoma: a mouse model of thyroid carcinogenesis. Thyroid. 2002;12:963–969. [PubMed] [Google Scholar]
104. Furuya F, Hanover JA, Cheng SY. Activation of phosphatidylinositol 3-kinase signaling by a mutant thyroid hormone β receptor. Proc Natl Acad Sci USA. 2006;103:1780–1785. This study demonstrates robust activation of the PI3K–AKT pathway by a mutant thyroid hormone β-receptor that causes FTC in mice, thus providing the first animal data to support an important role of the PI3K–AKT pathway in the tumorigenesis of FTC. [PMC free article] [PubMed] [Google Scholar]
105. Kim CS, et al. AKT activation promotes metastasis in a mouse model of follicular thyroid carcinoma. Endocrinology. 2005;146:4456–4463. [PubMed] [Google Scholar]
106. Saji M, et al. Akt1 deficiency delays tumor progression, vascular invasion, and distant metastasis in a murine model of thyroid cancer. Oncogene. 2011;30:4307–4315. [PMC free article] [PubMed] [Google Scholar]
107. Xu PZ, Chen ML, Jeon SM, Peng XD, Hay N. The effect Akt2 deletion on tumor development in Pten+/− mice. Oncogene. 2012;31:518–526. [PMC free article] [PubMed] [Google Scholar]
108. Liu D, Hou P, Liu Z, Wu G, Xing M. Genetic alterations in the phosphoinositide 3-kinase/Akt signaling pathway confer sensitivity of thyroid cancer cells to therapeutic targeting of Akt and mammalian target of rapamycin. Cancer Res. 2009;69:7311–7319. This study demonstrates that the efficacy of PI3K– AKT inhibitors in thyroid cancer cells depends on genetic alterations in the pathway, thus providing the first preclinical evidence supporting a genetic-based therapeutic approach targeting the PI3K–AKT pathway in thyroid cancer. [PMC free article] [PubMed] [Google Scholar]
109. Liu R, et al. The Akt-specific inhibitor MK2206 selectively inhibits thyroid cancer cells harboring mutations that can activate the PI3K/Akt pathway. J Clin Endocrinol Metab. 2011;96:e577–e585. [PMC free article] [PubMed] [Google Scholar]
110. LiVolsi VA, Baloch ZW. Follicular-patterned tumors of the thyroid: the battle of benign versus malignant versus so-called uncertain. Endocr Pathol. 2011;22:184–189. This is an excellent discussion of the criteria and challenges for defining follicular thyroid tumours. [PubMed] [Google Scholar]
111. Abbosh PH, Nephew KP. Multiple signaling pathways converge on β-catenin in thyroid cancer. Thyroid. 2005;15:551–561. [PubMed] [Google Scholar]
112. Burrows N, et al. Expression of hypoxia-inducible factor 1α in thyroid carcinomas. Endocr Relat Cancer. 2010;17:61–72. [PMC free article] [PubMed] [Google Scholar]
113. Burrows N, et al. GDC-0941 inhibits metastatic characteristics of thyroid carcinomas by targeting both the phosphoinositide-3 kinase (PI3K) and hypoxia-inducible factor-1α (HIF-1α) pathways. J Clin Endocrinol Metab. 2011;96:e1934–e1943. [PubMed] [Google Scholar]
114. Guigon CJ, Zhao L, Willingham MC, Cheng SY. PTEN deficiency accelerates tumour progression in a mouse model of thyroid cancer. Oncogene. 2009;28:509–517. [PMC free article] [PubMed] [Google Scholar]
115. Karin M. NF-κB as a critical link between inflammation and cancer. Cold Spring Harb Perspect Biol. 2009;1:a000141. [PMC free article] [PubMed] [Google Scholar]
116. Pacifico F, et al. Oncogenic and anti-apoptotic activity of NF-κB in human thyroid carcinomas. J Biol Chem. 2004;279:54610–54619. [PubMed] [Google Scholar]
117. Visconti R, et al. Expression of the neoplastic phenotype by human thyroid carcinoma cell lines requires NFκB p65 protein expression. Oncogene. 1997;15:1987–1994. [PubMed] [Google Scholar]
118. Mitsiades CS, et al. Antitumor effects of the proteasome inhibitor bortezomib in medullary and anaplastic thyroid carcinoma cells in vitro. J Clin Endocrinol Metab. 2006;91:4013–4021. [PubMed] [Google Scholar]
119. Pacifico F, Leonardi A. Role of NF-κB in thyroid cancer. Mol Cell Endocrinol. 2010;321:29–35. [PubMed] [Google Scholar]
120. Li X, Abdel-Mageed AB, Mondal D, Kandil E. The nuclear factor κ-B signaling pathway as a therapeutic target against thyroid cancers. Thyroid. 2012 Dec 28; doi: 10.1089/thy.2012.0237. [PubMed] [CrossRef] [Google Scholar]
121. Ikenoue T, et al. Functional analysis of mutations within the kinase activation segment of B-Raf in human colorectal tumors. Cancer Res. 2003;63:8132–8137. [PubMed] [Google Scholar]
122. Ikenoue T, et al. Different effects of point mutations within the B-Raf glycine-rich loop in colorectal tumors on mitogen-activated protein/extracellular signal-regulated kinase kinase/extracellular signal-regulated kinase and nuclear factor κB pathway and cellular transformation. Cancer Res. 2004;64:3428–3435. [PubMed] [Google Scholar]
123. Vale T, Ngo TT, White MA, Lipsky PE. Raf-induced transformation requires an interleukin 1 autocrine loop. Cancer Res. 2001;61:602–607. [PubMed] [Google Scholar]
124. Liu D, Xing M. Potent inhibition of thyroid cancer cells by the MEK inhibitor PD0325901 and its potentiation by suppression of the PI3K and NF-κB pathways. Thyroid. 2008;18:853–864. [PMC free article] [PubMed] [Google Scholar]
125. Oh HJ, et al. Role of the tumor suppressor RASSF1A in Mst1-mediated apoptosis. Cancer Res. 2006;66:2562–2569. [PubMed] [Google Scholar]
126. Lehtinen MK, et al. A conserved MST-FOXO signaling pathway mediates oxidative-stress responses and extends life span. Cell. 2006;125:987–1001. [PubMed] [Google Scholar]
127. Schagdarsurengin U, et al. Frequent epigenetic silencing of the CpG island promoter of RASSF1A in thyroid carcinoma. Cancer Res. 2002;62:3698–3701. [PubMed] [Google Scholar]
128. Xing M, et al. Early occurrence of RASSF1A hypermethylation and its mutual exclusion with BRAF mutation in thyroid tumorigenesis. Cancer Res. 2004;64:1664–1668. This is the first study to link BRAF-V600E to the RASSF1 pathway in thyroid cancer. [PubMed] [Google Scholar]
129. Van Der Heide LP, Hoekman MF, Smidt MP. The ins and outs of FoxO shuttling: mechanisms of FoxO translocation and transcriptional regulation. Biochem J. 2004;380:297–309. [PMC free article] [PubMed] [Google Scholar]
130. Clevers H, Nusse R. Wnt/β-catenin signaling and disease. Cell. 2012;149:1192–1205. [PubMed] [Google Scholar]
131. Wiseman SM, Griffith OL, Gown A, Walker B, Jones SJ. Immunophenotyping of thyroid tumors identifies molecular markers altered during transformation of differentiated into anaplastic carcinoma. Am J Surg. 2011;201:580–586. [PubMed] [Google Scholar]
132. Sastre-Perona A, Santisteban P. Role of the wnt pathway in thyroid cancer. Front Endocrinol. 2012;3:31. [PMC free article] [PubMed] [Google Scholar]
133. Lu C, Zhu X, Willingham MC, Cheng SY. Activation of tumor cell proliferation by thyroid hormone in a mouse model of follicular thyroid carcinoma. Oncogene. 2012;31:2007–2016. [PMC free article] [PubMed] [Google Scholar]
134. Cross DA, Alessi DR, Cohen P, Andjelkovich M, Hemmings BA. Inhibition of glycogen synthase kinase-3 by insulin mediated by protein kinase B. Nature. 1995;378:785–789. [PubMed] [Google Scholar]
135. Choi KS, Bae MK, Jeong JW, Moon HE, Kim KW. Hypoxia-induced angiogenesis during carcinogenesis. J Biochem Mol Biol. 2003;36:120–127. [PubMed] [Google Scholar]
136. Scarpino S, et al. Increased expression of Met protein is associated with up-regulation of hypoxia inducible factor-1 (HIF-1) in tumour cells in papillary carcinoma of the thyroid. J Pathol. 2004;202:352–358. [PubMed] [Google Scholar]
137. Kimura T, et al. Regulation of thyroid cell proliferation by TSH and other factors: a critical evaluation of in vitro models. Endocr Rev. 2001;22:631–656. [PubMed] [Google Scholar]
138. Lu C, Zhao L, Ying H, Willingham MC, Cheng SY. Growth activation alone is not sufficient to cause metastatic thyroid cancer in a mouse model of follicular thyroid carcinoma. Endocrinology. 2010;151:1929–1939. This study demonstrated for the first time that a transgenic mouse model of thyroid tumorigenesis promoted by a mutation (in this case, mutant THRβ) requires an intact TSHR pathway. [PMC free article] [PubMed] [Google Scholar]
139. Franco AT, et al. Thyrotrophin receptor signaling dependence of Braf-induced thyroid tumor initiation in mice. Proc Natl Acad Sci USA. 2011;108:1615–1620. Similarly to reference 138, this study also demonstrated a requirement for an intact TSHR pathway, this time for BRAF-V600E-induced thyroid tumorigenesis in transgenic mice. [PMC free article] [PubMed] [Google Scholar]
140. Boelaert K, et al. Serum thyrotropin concentration as a novel predictor of malignancy in thyroid nodules investigated by fine-needle aspiration. J Clin Endocrinol Metab. 2006;91:4295–4301. [PubMed] [Google Scholar]
141. Fiore E, et al. Lower levels of TSH are associated with a lower risk of papillary thyroid cancer in patients with thyroid nodular disease: thyroid autonomy may play a protective role. Endocr Relat Cancer. 2009;16:1251–1260. [PubMed] [Google Scholar]
142. Haymart MR, et al. Higher serum thyroid stimulating hormone level in thyroid nodule patients is associated with greater risks of differentiated thyroid cancer and advanced tumor stage. J Clin Endocrinol Metab. 2008;93:809–814. [PMC free article] [PubMed] [Google Scholar]
143. Lyons J, et al. Two G protein oncogenes in human endocrine tumors. Science. 1990;249:655–659. [PubMed] [Google Scholar]
144. Parma J, et al. Somatic mutations in the thyrotropin receptor gene cause hyperfunctioning thyroid adenomas. Nature. 1993;365:649–651. [PubMed] [Google Scholar]
145. Gudmundsson J, et al. Common variants on 9q22.33 and 14q13.3 predispose to thyroid cancer in European populations. Nature Genet. 2009;41:460–464. This is the first large genome-wide association study to reveal several novel single-nucleotide polymorphisms that are associated with an increased risk of thyroid cancer and low levels of serum TSH. [PMC free article] [PubMed] [Google Scholar]
146. Gudmundsson J, et al. Discovery of common variants associated with low TSH levels and thyroid cancer risk. Nature Genet. 2012;44:319–322. [PMC free article] [PubMed] [Google Scholar]
147. Wu G, et al. Uncommon mutation, but common amplifications, of the PIK3CA gene in thyroid tumors. J Clin Endocrinol Metab. 2005;90:4688–4693. [PubMed] [Google Scholar]
148. Costa AM, et al. BRAF mutation associated with other genetic events identifies a subset of aggressive papillary thyroid carcinoma. Clin Endocrinol. 2008;68:618–634. [PubMed] [Google Scholar]
149. Henderson YC, et al. High rate of BRAF and RET/PTC dual mutations associated with recurrent papillary thyroid carcinoma. Clin Cancer Res. 2009;15:485–491. [PMC free article] [PubMed] [Google Scholar]
150. Dankort D, et al. BrafV600E cooperates with Pten loss to induce metastatic melanoma. Nature Genet. 2009;41:544–552. [PMC free article] [PubMed] [Google Scholar]
151. Riesco-Eizaguirre G, Santisteban P. New insights in thyroid follicular cell biology and its impact in thyroid cancer therapy. Endocr Relat Cancer. 2007;14:957–977. [PubMed] [Google Scholar]
152. Jhiang SM. Regulation of sodium/iodide symporter. Rev Endocr Metab Disord. 2000;1:205–215. [PubMed] [Google Scholar]
153. Riesco-Eizaguirre G, Gutierrez-Martinez P, Garcia-Cabezas MA, Nistal M, Santisteban P. The oncogene BRAF V600E is associated with a high risk of recurrence and less differentiated papillary thyroid carcinoma due to the impairment of Na+/I targeting to the membrane. Endocr Relat Cancer. 2006;13:257–269. [PubMed] [Google Scholar]
154. Mian C, et al. Molecular characteristics in papillary thyroid cancers (PTCs) with no 131I uptake. Clin Endocrinol. 2008;68:108–116. [PubMed] [Google Scholar]
155. Barollo S, et al. BRAF in primary and recurrent papillary thyroid cancers: the relationship with 131I and 2-[18F]fluoro-2-deoxy-D-glucose uptake ability. Eur J Endocrinol. 2010;163:659–663. [PubMed] [Google Scholar]
156. Durante C, et al. BRAF mutations in papillary thyroid carcinomas inhibit genes involved in iodine metabolism. J Clin Endocrinol Metab. 2007;92:2840– 2843. [PubMed] [Google Scholar]
157. Romei C, et al. BRAFV600E mutation, but not RET/ PTC rearrangements, is correlated with a lower expression of both thyroperoxidase and sodium iodide symporter genes in papillary thyroid cancer. Endocr Relat Cancer. 2008;15:511–520. [PubMed] [Google Scholar]
158. Oler G, Cerutti JM. High prevalence of BRAF mutation in a Brazilian cohort of patients with sporadic papillary thyroid carcinomas: correlation with more aggressive phenotype and decreased expression of iodide-metabolizing genes. Cancer. 2009;115:972–980. [PubMed] [Google Scholar]
159. Espadinha C, Santos JR, Sobrinho LG, Bugalho MJ. Expression of iodine metabolism genes in human thyroid tissues: evidence for age and BRAFV600E mutation dependency. Clin Endocrinol. 2009;70:629–635. [PubMed] [Google Scholar]
160. Morari EC, et al. Use of sodium iodide symporter expression in differentiated thyroid carcinomas. Clin Endocrinol. 2011;75:247–254. [PubMed] [Google Scholar]
161. Liu D, et al. Suppression of BRAF/MEK/MAP kinase pathway restores expression of iodide-metabolizing genes in thyroid cells expressing the V600E BRAF mutant. Clin Cancer Res. 2007;13:1341–1349. This study for the first time directly demonstrated in an in vitro thyroid cell system that BRAF-V600E silences thyroid iodide-handling genes and that suppression of BRAF-V600E or MEK can restore the expression of these genes, which was later reproduced in transgenic mouse studies in reference 163. [PubMed] [Google Scholar]
162. Vadysirisack DD, Venkateswaran A, Zhang Z, Jhiang SM. MEK signaling modulates sodium iodide symporter at multiple levels and in a paradoxical manner. Endocr Relat Cancer. 2007;14:421–432. [PubMed] [Google Scholar]
163. Chakravarty D, et al. Small-molecule MAPK inhibitors restore radioiodine incorporation in mouse thyroid cancers with conditional BRAF activation. J Clin Invest. 2011;121:4700–4711. [PMC free article] [PubMed] [Google Scholar]
164. Hou P, Bojdani E, Xing M. Induction of thyroid gene expression and radioiodine uptake in thyroid cancer cells by targeting major signaling pathways. J Clin Endocrinol Metab. 2010;95:820–828. This study demonstrated robust expression of thyroid iodide-handling genes and radioiodine uptake in thyroid cancer cells by simultaneously targeting MAPK, PI3K–AKT and histone deacetylase systems, which could be enhanced by TSH, thus providing strong clinical implications for the treatment of thyroid cancer. [PMC free article] [PubMed] [Google Scholar]
165. Liu Z, Xing M. Induction of sodium/iodide symporter (NIS) expression and radioiodine uptake in non-thyroid cancer cells. PLoS ONE. 2012;7:e31729. [PMC free article] [PubMed] [Google Scholar]
166. Costamagna E, Garcia B, Santisteban P. The functional interaction between the paired domain transcription factor Pax8 and Smad3 is involved in transforming growth factor-β repression of the sodium/ iodide symporter gene. J Biol Chem. 2004;279:3439–3446. [PubMed] [Google Scholar]
167. de Souza EC, et al. MTOR downregulates iodide uptake in thyrocytes. J Endocrinol. 2010;206:113–120. [PubMed] [Google Scholar]
168. Garcia B, Santisteban P. PI3K is involved in the IGF-I inhibition of TSH-induced sodium/iodide symporter gene expression. Mol Endocrinol. 2002;16:342–352. [PubMed] [Google Scholar]
169. Kogai T, Sajid-Crockett S, Newmarch LS, Liu YY, Brent GA. Phosphoinositide-3-kinase inhibition induces sodium/iodide symporter expression in rat thyroid cells and human papillary thyroid cancer cells. J Endocrinol. 2008;199:243–252. [PubMed] [Google Scholar]
170. Xing M, Haugen BR, Schlumberger M. Promises in molecular-based management of differentiated thyroid cancer. Lancet. (in the press). This is a review on the clinically promising diagnostic and prognostic molecular markers and molecular therapeutic targets in thyroid cancer. It summarizes the recent developments in clinical translational research of thyroid cancer medicine. [Google Scholar]
171. Cohen Y, et al. Mutational analysis of BRAF in fine needle aspiration biopsies of the thyroid: a potential application for the preoperative assessment of thyroid nodules. Clin Cancer Res. 2004;10:2761–2765. [PubMed] [Google Scholar]
172. Xing M, et al. Detection of BRAF mutation on fine needle aspiration biopsy specimens: a new diagnostic tool for papillary thyroid cancer. J Clin Endocrinol Metab. 2004;89:2867–2872. References 171 and 172 are the first studies to demonstrate the feasibility of BRAF mutation testing on thyroid FNAB samples for the diagnosis of thyroid cancer. [PubMed] [Google Scholar]
173. Xing M. Prognostic utility of BRAF mutation in papillary thyroid cancer. Mol Cell Endocrinol. 2010;321:86–93. [PMC free article] [PubMed] [Google Scholar]
174. Filicori F, et al. Risk stratification of indeterminate thyroid fine-needle aspiration biopsy specimens based on mutation analysis. Surgery. 2011;150:1085–1091. [PubMed] [Google Scholar]
175. Nikiforov YE, et al. Impact of mutational testing on the diagnosis and management of patients with cytologically indeterminate thyroid nodules: a prospective analysis of 1056 FNA samples. J Clin Endocrinol Metab. 2011;96:3390–3397. This is a large study of the mutational analyses of thyroid nodules for their diagnostic utility. The analyses show high diagnostic specificity and positive predictive value. [PMC free article] [PubMed] [Google Scholar]
176. Alexander EK, et al. Preoperative diagnosis of benign thyroid nodules with indeterminate cytology. N Engl J Med. 2012;367:705–715. This study demonstrates a high sensitivity and negative predictive power of a gene expression classifier for diagnosing benign thyroid tumours with indeterminate cytology. [PubMed] [Google Scholar]
177. Xing M, et al. The BRAF T1799A mutation and poor outcomes of papillary thyroid cancer-report from the international collaborative BRAF study group. Thyroid. 2011;21(S1):Abstract 112. [Google Scholar]
178. Kebebew E, et al. The prevalence and prognostic value of BRAF mutation in thyroid cancer. Ann Surg. 2007;246:466–470. [PMC free article] [PubMed] [Google Scholar]
179. Basolo F, et al. Correlation between the BRAF V600E mutation and tumor invasiveness in papillary thyroid carcinomas smaller than 20 millimeters: analysis of 1060 cases. J Clin Endocrinol Metab. 2010;95:4197–4205. [PubMed] [Google Scholar]
180. Lin KL, et al. The BRAF mutation is predictive of aggressive clinicopathological characteristics in papillary thyroid microcarcinoma. Ann Surg Oncol. 2010;17:3294–3300. [PubMed] [Google Scholar]
181. Elisei R, et al. The BRAFV600E mutation is an independent, poor prognostic factor for the outcome of patients with low-risk intrathyroid papillary thyroid carcinoma: single-institution results from a large cohort study. J Clin Endocrinol Metab. 2012;97:4390–4398. [PubMed] [Google Scholar]
182. Xing M, et al. BRAF mutation testing of thyroid fine-needle aspiration biopsy specimens for preoperative risk stratification in papillary thyroid cancer. J Clin Oncol. 2009;27:2977–2982. [PMC free article] [PubMed] [Google Scholar]
183. Joo JY, et al. Prediction of occult central lymph node metastasis in papillary thyroid carcinoma by preoperative BRAF analysis using fine-needle aspiration biopsy: a prospective study. J Clin Endocrinol Metab. 2012;97:3996–4003. [PubMed] [Google Scholar]
184. Garcia-Rostan G, et al. ras mutations are associated with aggressive tumor phenotypes and poor prognosis in thyroid cancer. J Clin Oncol. 2003;21:3226–3235. [PubMed] [Google Scholar]
185. Basolo F, et al. N-ras mutation in poorly differentiated thyroid carcinomas: correlation with bone metastases and inverse correlation to thyroglobulin expression. Thyroid. 2000;10:19–23. [PubMed] [Google Scholar]
186. Volante M, et al. RAS mutations are the predominant molecular alteration in poorly differentiated thyroid carcinomas and bear prognostic impact. J Clin Endocrinol Metab. 2009;94:4735–4741. [PubMed] [Google Scholar]
187. Fukahori M, et al. The association between RAS gene mutations and clinical characteristics in follicular thyroid tumors: new insights from a single center and a large patient cohort. Thyroid. 2012;22:683–689. [PubMed] [Google Scholar]
188. Cohen EE, et al. Axitinib is an active treatment for all histologic subtypes of advanced thyroid cancer: results from a phase II study. J Clin Oncol. 2008;26:4708–4713. [PMC free article] [PubMed] [Google Scholar]
189. Gupta-Abramson V, et al. Phase II trial of sorafenib in advanced thyroid cancer. J Clin Oncol. 2008;26:4714–4719. [PMC free article] [PubMed] [Google Scholar]
190. Kloos RT, et al. Phase II trial of sorafenib in metastatic thyroid cancer. J Clin Oncol. 2009;27:1675–1684. [PMC free article] [PubMed] [Google Scholar]
191. Ahmed M, et al. Analysis of the efficacy and toxicity of sorafenib in thyroid cancer: a phase II study in a UK based population. Eur J Endocrinol. 2011;165:315–322. [PubMed] [Google Scholar]
192. Sherman SI, et al. Motesanib diphosphate in progressive differentiated thyroid cancer. N Engl J Med. 2008;359:31–42. This study represents an outstanding example of the success of recent clinical trials of novel small-molecule inhibitors targeting aberrantly activated signalling pathways in thyroid cancer. [PubMed] [Google Scholar]
193. Bible KC, et al. Efficacy of pazopanib in progressive, radioiodine-refractory, metastatic differentiated thyroid cancers: results of a phase 2 consortium study. Lancet Oncol. 2010;11:962–972. [PMC free article] [PubMed] [Google Scholar]
194. Xing M. Genetic-targeted therapy of thyroid cancer: a real promise. Thyroid. 2009;19:805–809. [PubMed] [Google Scholar]
195. Ball DW, et al. Selective growth inhibition in BRAF mutant thyroid cancer by the MEK 1/2 inhibitor AZD6244 (ARRY-142886) J Clin Endocrinol Metab. 2007;92:4712–4718. [PubMed] [Google Scholar]
196. Leboeuf R, et al. BRAFV600E mutation is associated with preferential sensitivity to mitogen-activated protein kinase kinase inhibition in thyroid cancer cell lines. J Clin Endocrinol Metab. 2008;93:2194–2201. [PMC free article] [PubMed] [Google Scholar]
197. Liu D, Liu Z, Jiang D, Dackiw AP, Xing M. Inhibitory effects of the MEK inhibitor CI-1040 on the proliferation and tumor growth of thyroid cancer cells with BRAF or RAS mutations. J Clin Endocrinol Metab. 2007;92:4686–4695. [PubMed] [Google Scholar]
198. Schweppe RE, et al. Distinct genetic alterations in the mitogen-activated protein kinase pathway dictate sensitivity of thyroid cancer cells to mitogen-activated protein kinase kinase 1/2 inhibition. Thyroid. 2009;19:825–835. [PMC free article] [PubMed] [Google Scholar]
199. Liu D, Xing J, Trink B, Xing M. BRAF mutation-selective inhibition of thyroid cancer cells by the novel MEK inhibitor RDEA119 and genetic-potentiated synergism with the mTOR inhibitor temsirolimus. Int J Cancer. 2010;127:2965–2973. [PMC free article] [PubMed] [Google Scholar]
200. Salerno P, et al. Cytostatic activity of adenosine triphosphate-competitive kinase inhibitors in BRAF mutant thyroid carcinoma cells. J Clin Endocrinol Metab. 2010;95:450–455. [PubMed] [Google Scholar]
201. Xing J, Liu R, Xing M, Trink B. The BRAFT1799A mutation confers sensitivity of thyroid cancer cells to the BRAFV600E inhibitor PLX4032 (RG7204) Biochem Biophys Res Commun. 2011;404:958–962. [PMC free article] [PubMed] [Google Scholar]
202. Flaherty KT, et al. Inhibition of mutated, activated BRAF in metastatic melanoma. N Engl J Med. 2010;363:809–819. [PMC free article] [PubMed] [Google Scholar]
203. Janku F, et al. PI3K/AKT/mTOR inhibitors in patients with breast and gynecologic malignancies harboring PIK3CA mutations. J Clin Oncol. 2012;30:777–782. [PMC free article] [PubMed] [Google Scholar]
204. Jin N, Jiang T, Rosen DM, Nelkin BD, Ball DW. Synergistic action of a RAF inhibitor and a dual PI3K/mTOR inhibitor in thyroid cancer. Clin Cancer Res. 2011;17:6482–6489. [PMC free article] [PubMed] [Google Scholar]
205. Liu R, Liu D, Xing M. The Akt inhibitor MK2206 synergizes, but perifosine antagonizes, the BRAF(V600E) inhibitor PLX4032 and the MEK1/2 inhibitor AZD6244 in the inhibition of thyroid cancer cells. J Clin Endocrinol Metab. 2012;97:e173–e182. [PMC free article] [PubMed] [Google Scholar]
206. Vasko V, et al. Specific pattern of RAS oncogene mutations in follicular thyroid tumors. J Clin Endocrinol Metab. 2003;88:2745–2752. [PubMed] [Google Scholar]
207. Nikiforova MN, et al. RAS point mutations and PAX8-PPARγ rearrangement in thyroid tumors: evidence for distinct molecular pathways in thyroid follicular carcinoma. J Clin Endocrinol Metab. 2003;88:2318–2326. [PubMed] [Google Scholar]
208. Rivera M, et al. Molecular genotyping of papillary thyroid carcinoma follicular variant according to its histological subtypes (encapsulated versus infiltrative) reveals distinct BRAF and RAS mutation patterns. Mod Pathol. 2010;23:1191–1200. [PMC free article] [PubMed] [Google Scholar]
209. Manenti G, Pilotti S, Re FC, Della PG, Pierotti MA. Selective activation of ras oncogenes in follicular and undifferentiated thyroid carcinomas. Eur J Cancer. 1994;30A:987–993. [PubMed] [Google Scholar]
210. Zhu Z, Gandhi M, Nikiforova MN, Fischer AH, Nikiforov YE. Molecular profile and clinicalpathologic features of the follicular variant of papillary thyroid carcinoma. An unusually high prevalence of ras mutations. Am J Clin Pathol. 2003;120:71–77. [PubMed] [Google Scholar]
211. Santarpia L, et al. Genetic alterations in the RAS/ RAF/mitogen-activated protein kinase and phosphatidylinositol 3-kinase/Akt signaling pathways in the follicular variant of papillary thyroid carcinoma. Cancer. 2010;116:2974–2983. [PubMed] [Google Scholar]
212. Dahia PL, et al. Somatic deletions and mutations in the Cowden disease gene, PTEN, in sporadic thyroid tumors. Cancer Res. 1997;57:4710–4713. [PubMed] [Google Scholar]
-