Skip to main content

Genetic susceptibility to hereditary non-medullary thyroid cancer

Abstract

Non-medullary thyroid cancer (NMTC) is the most common type of thyroid cancer. With the increasing incidence of NMTC in recent years, the familial form of the disease has also become more common than previously reported, accounting for 5–15% of NMTC cases. Familial NMTC is further classified as non-syndromic and the less common syndromic FNMTC. Although syndromic NMTC has well-known genetic risk factors, the gene(s) responsible for the vast majority of non-syndromic FNMTC cases are yet to be identified. To date, several candidate genes have been identified as susceptibility genes in hereditary NMTC. This review summarizes genetic predisposition to non-medullary thyroid cancer and expands on the role of genetic variants in thyroid cancer tumorigenesis and the level of penetrance of NMTC-susceptibility genes.

Peer Review reports

Introduction

Thyroid cancer is the most common endocrine malignancy [1], with its global incidence rate increasing substantially in the past four decades [2]. Thyroid cancers can originate due to the accumulation of genetic mutations in para-follicular or follicular cells. Thyroid cancers originating from para-follicular calcitonin-producing C cells are known as medullary thyroid carcinoma (MTC) and account for 5% of all cases, whereas the more common type of thyroid cancer arises from follicular cells and is known as non-medullary thyroid cancer [1]. The majority of NMTC are differentiated thyroid cancers (DTC) which include papillary and follicular thyroid cancers. Papillary thyroid cancer (PTC) accounts for more than 85% of NMTC cases and follicular thyroid cancer (FTC) accounts for 10–15% of NMTC cases. The rare forms of NMTC are poorly differentiated thyroid carcinomas and anaplastic thyroid carcinomas [3, 4]. Over 90% of all thyroid cancers are sporadic and arise from somatic genetic changes [5]. The remaining are familial forms of NMTC and MTC. Familial MTC (FMTC) has well-known genetic alterations and genotype-phenotype correlations. On the contrary, the genetic causes of familial NMTC (FNMTC), or familial follicular cell-derived carcinoma are poorly understood [6, 7]. FNMTC is clinically defined as the presence of the disease in two or more first-degree relatives of the patient. FNMTC can further be classified as syndromic or non-syndromic, depending on whether the thyroid cancer is the primary cancer (non-syndromic) or as a part of one of many constellations of tumours in kindreds (syndromic FNMTC) [8]. Hereditary cancer syndromes associated with FNMTC account for 5% of all familial cases and include Familial adenomatous polyposis, Cowden syndrome, Carney complex, Werner syndrome, DICER1 syndrome, Ataxia-telangiectasia, Bannayan-Riley-Ruvalcaba syndrome, Li-Fraumeni syndrome, Peutz-Jeghers syndrome, and Pendred syndrome  (Fig. 1) [9].

Fig. 1
figure 1

Overview of thyroid cancer subtype classification

According to the Swedish family cancer database, the proportion of cancer susceptibility accounted for by genetic factors was the highest for thyroid cancer among 15 cancer sites [10]. Additionally, family and twin studies from Utah and Sweden suggested thyroid cancer as one of the most heritable cancers displaying Mendelian inheritance, with reported risks of 8- to 12-fold higher for first-degree relatives of thyroid cancer patients compared to the general population [11,12,13] . A family cohort of the Norwegian cancer registry database estimated that the familial risk ratio of NMTC in affected first-degree relatives is 5.2 for men and 4.9 for women [14, 15]. Similarly, Lin et al. identified the family structures of 38,686 NMTC patients in Taiwan. The prevalence of NMTC in the general population and in first-degree relatives of NMTC patients were 0.16% and 0.64%, respectively. This shows a 5.47-fold increased risk for NMTC for first-degree family members [16]. FNMTC patients present with more aggressive disease at a younger age compared to sporadic cases, this includes larger tumours with more lymph node involvement [17,18,19]. However, studies have shown no significant increase in risk of recurrence or disease-related mortality in FNMTC cases compared to sporadic cases [20,21,22]. As a result, the evidence for a worst disease outcome in FNMTC compared to sporadic cases is conflicting [18, 23]. Additionally, the second generation of FNMTC patients present at a younger age with more severe symptoms, indicating the presence of clinical anticipation [24]. In a prospective cohort study, at-risk relatives of twenty-five kindreds with two or more members affected by FNMTC were screened with neck ultrasound and fine-needle aspiration biopsy of thyroid nodules. The results indicated the presence of thyroid cancer in 4.6% of families with two affected members and 22.7% of families with three or more affected members. The tumours that were identified with screening were smaller in size, had less lymph node metastases, and required less extensive treatments. Therefore, the early detection of FNMTC can potentially improve the treatment outcome [25].

Despite the solid evidence for the heritability of thyroid cancer, only a handful of variants have been convincingly associated with a higher risk of this cancer. The high heritability of thyroid cancer is likely due to the contributions of rare but high-penetrance mutations in some cases or common but low-penetrance variants in others [26]. The present study aimed to summarize the literature regarding variants associated with higher risk of hereditary NMTC and provide a more extensive background on the penetrance, molecular function and functional consequences of these mutations, which can further clarify the etiology of thyroid cancer and aid in the identification of disease risk in family members of NMTC patients.

Genetic variants associated with risk of non-medullary thyroid cancer

Approximately 5–15% of NMTC cases occur due to germline mutations [17]. Genetic variants conferring risk of complex disorders such as cancer are either rare mutations with moderate to high penetrance or common variants with low penetrance. The genetic predisposition to NMTC seems to be relatively strong based on the previous case-control studies. To date, multiple susceptibility genes have been identified through genome-wide association studies (GWAS) (Table 1). Methods such as family-based exome sequencing, next generation sequencing and linkage studies also identified several susceptibility loci associated with NMTC. Genetic variants from these genes have been classified based on their level of penetrance using odds ratio (OR) compiled from previous case-control studies. In this review, variants with an OR lower than 1.5 were classified as low penetrance mutations, and those with an OR between 1.5 and 2.5 were classified as moderate penetrance and highly penetrant mutations were categorized as variants with an OR greater than 2.5 [34].

Table 1 Variants associated with increased risk for non-medullary thyroid cancer in various populations identified by GWAS

Moderate and high penetrant mutations

FOXE1

The FOXE1 (forkhead factor E1) gene is located at chromosome 9q22.33 and encodes for the FOXE1 transcription factor (thyroid transcription factor 2, TTF-2), which regulates thyroglobulin and thyroperoxidase gene expression. In a genome-wide association study in a population of 192 and 37,196 thyroid cancer cases and controls, seven of the nine strongest association signals were in a similar linkage disequilibrium region as the FOXE1 gene. Further replication results from 241 patients in combination with results from a GWAS showed the strongest association signal for allele A of rs965513 with an OR of 1.75 in European populations from Iceland, Columbus, and Spain [35]. In a study by Landa et al. (2009), another single nucleotide polymorphism (SNP) (rs1867277) located in the 5’UTR region of FOXE1 gene was positively associated with thyroid cancer in Spanish and Italian cohorts. The authors proposed this variant as a causal SNP in susceptibility to thyroid cancer through DNA binding assays and transfection studies. The variant was observed to cause cancer susceptibility through the recruitment of USF1/USF2 transcription factors [36].

Furthermore, the first study between FOXE1 gene and susceptibility to FNMTC was reported in 2012. Nine exonic and promoter variants of FOXE1 gene were studied in a population of 60 Portuguese FNMTC probands and 80 sporadic cases with matched controls (Table 2). As a result, rs965513 and rs1867277 were associated with increased risk of FNMTC. The authors also observed an association between FOXE1 polyalanine tract expansions and familial thyroid cancer risk (OR = 2.56) [37]. The same group identified a rare FOXE1 variant (p.A248G) which co-segregated with thyroid cancer in one family and was also present in a case of sporadic NMTC. Further In vitro studies showed that this variant promoted cell migration and proliferation [38]. A large family-based study with 672 subjects belonging to 133 pedigrees with FNMTC cases genotyped twenty-three variants on 11 loci. Only three variants of 9q22.33 near FOXE1 showed a positive association with FNMTC. FOXE1 gene variant rs1867277 had an OR of 3.17 under a recessive mode of inheritance. The other two variants also showed high penetrance under the recessive model (OR = 4.63 for rs10759944 and OR = 5.10 for rs965513) [39]. The rs965513 variant has been previously associated with increased tumour size and extrathyroidal expansions in PTC patients [40]. In another two-step association study involving 1820 DTC cases and 2410 controls in Europe, two moderate penetrant FOXE1 variants were identified, rs7028661 with an OR of 1.64 and rs7037324 with an OR of 1.54 [33]. He et al. showed that the rs965513 variant and 4 other variants in close proximity regulate FOXE1 and the PTC susceptibility candidate 2 (PTCSC2) gene transcriptional activity through regulatory enhancers [41]. PTCSC2 is novel long non-coding RNA (lncRNA) gene with its transcripts downregulated in PTC tumours. Later, myosin-9 (MYH9) was identified as a PTCSC2 binding protein with the ability to inhibit the promoter shared by FOXE1 and PTCSC2 in both directions. Thus, PTC risk is potentially conferred by the interaction between a lncRNA (PTCSC2), MYH9, and FOXE1 [42].

Table 2 FOXE1 variants associated with hereditary thyroid cancer

HABP2

The Hyaluronan-Binding Protein 2 (HABP2) gene is located on chromosome 10q25.3 and encodes a member of the peptidase S1 family of serine proteases [43]. Mutations in this gene are associated with non-medullary thyroid cancer and susceptibility to venous thromboembolism. Gara et al. (2015) performed whole-exome sequencing of peripheral blood in 7 affected members of a FNMTC kindred and unaffected spouses as controls, a germline variant (G534E; rs7080536) was identified in the HABP2 gene. All affected family members were heterozygous for the variant in peripheral blood DNA. This mutation was found with an allele frequency of 2.2% in the ExAC database, and in 4.7% of 423 patients with sporadic thyroid cancer reported in Human Cancer Genome Atlas multiethnic database. Functional studies confirmed this loss of function variant’s pathogenicity and showed normal HABP2 has tumour-suppressive functionality [43]. Zhou et al. [44], Sponziello et al. [45], and Tomsic et al. [46] later argued that the allele frequency of the G534E variant exceeds the filtering criterion used by Gara et al. (less than 1% in public databases) and the role of this variant in thyroid cancer requires more studies. A replication study by Zhang et al. with a cohort of 64 subjects from 29 kindred, identified G534E variant in 6 PTC patients from 4 independent kindreds. The prevalence rate of 13.8% was reported for this variant in the 29 kindreds, suggesting HABP2 as a susceptibility gene for hereditary thyroid cancer [47]. However, NMTC risk conferred by HABP2 G534E was not confirmed by an association study of over 2000 NMTC cases and over 5000 population controls from the British Isles. The frequency of HABP2 G534 variant was 4.2% in cases and 4.6% in controls (OR = 0.74; P = 0.017) [48]. The data from various ethnic populations with large sample sizes suggest that this variant is unlikely to be a moderate or high penetrance gene in NMTC patients. Multiple other groups were unable to verify an association between the G534E variant and hereditary thyroid cancer [49,50,51,52,53,54,55,56,57]. Additionally, targeted sequencing of 516 PTC cases failed to identify the G534E variant. However, three other HABP2 variants (rs138864377, rs2286742, and rs3740530) were identified that can potentially increase the risk of PTC. The rs2286742 and rs3740530 variants in HABP2 had odds ratio of 9.644 and 3.989 in a recessive model, respectively (Table 3) [56]. However, no replication studies have been performed to identify the pathogenicity of these three HABP2 novel variants in PTC.

Table 3 HABP2 variants associated with hereditary thyroid cancer

Low penetrance mutations

TITF1/ NKX2.1 and PTCSC3

TITF1/NKX2.1 consists of two exons that encode thyroid-specific transcription factor-1 (TTF-1). A GWAS in the Icelandic population by Gudmundsson et al. identified an association between intergenic variant (rs944289) on chromosome 14q13.3 and risk for DTC (OR = 1.37) (Table 4) [35]. The closest gene to this variant is NKX2.1 [58]. This association was also confirmed by a case-control study in 1085 Korean DTC cases and 8884 controls that yielded an OR of 1.23 for rs944289 and an OR of 1.25 for the rs34081947 variant [29]. Another case-control study from the northern Chinese Han population also found an association between the rs944289 variant and PTC risk (OR = 1.23) [59]. On the contrary, a large family-based study of 672 subjects from 133 pedigree was not able to find any association between familial NMTC and rs944289 [39]. Studies on the rs944289 variant in different populations are summarized in Table 4. Ngan et al. performed targeted DNA sequencing for germline mutations in TITF-1/NKX2.1 in 20 patients with multinodular goiter (MNG) and PTC, 284 with only PTC, and 349 controls. In 4 of 20 unrelated patients with MNG/PTC, a germline mutation (A339V) was identified in NKX2.1/TITF-1. Only two of these 4 patients had a positive family history of PTC and the mutation showed an autosomal dominant pattern of inheritance. The mutation was not found among 349 healthy control subjects or among the 284 PTC patients who had no history of MNG [60]. In another study, none of the 63 familial PTC cases had the A339V mutation [61].

Table 4 TITF1/NKX2.1 variant rs944289 associated with hereditary thyroid cancer

Jendrzejewski et al. identified a non-coding RNA gene named papillary thyroid carcinoma susceptibility candidate 3 (PTCSC3) in a transcriptome gene expression analysis from 46 PTC tumour and unaffected thyroid tissue samples. Interestingly, PTCSC3 is located 3.2 kb downstream of rs944289 at 14q.13.3 and has lower expression in PTC thyroid tumours, suggesting a tumour-suppressor role. PTCS3 expression is reduced by the T allele of the rs944289 SNP which affects promoter activation. As a result, the risk allele of rs944289 can potentially decrease PTCSC3 promoter activation and thereby acts as a predisposition to PTC [62], [63].

SRGAP1

He et al. performed a genome-wide linkage analysis in 38 families with PTC and identified Slit-Robo GTPase-activating protein 1 (SRGAP1) as a candidate gene on chromosome 12q14.2. The SNPs, rs781626187 (Q149H) and rs797044990 (A275T) were two loss-of-function mutations in the Fes/CIP4 homology domain that segregated with PTC in one family each. Additionally, a missense variant (rs114817817) in the RhoGAP domain (R617C) also occurred in only one family [64]. The protein encoded by this gene is a GTPase activator and mutations in this gene can severely impair the ability to inactivate CDC42. CDC42 can mediate multiple signaling pathways, and plays a role in PTC tumourigenesis [65, 66]. To assess the frequency of the 4 missense variants in sporadic PTC cases and healthy controls, He et al. performed further targeted association studies on 2 large cohorts from Ohio and Poland which failed to confirm this association. In fact, Q149H and A275T were not found in 367 cases and 552 controls from Ohio or in the 432 cases and 424 controls from Poland. However, a SNP (rs2168411) located in intron 4 of SRGAP1 showed an association with PTC in both Ohio and Poland cohorts with a combined OR of 1.21 (95% CI 1.08–1.35, P = .0008). The rs114817817 variant was also identified in 4 of 742 sporadic cases of PTC in Ohio but in none of the 828 controls, which is suggestive of low penetrance. Future replication studies are required to confirm the candidacy of this variant [64].

NRG1

Previously, a SNP (rs2439302) on chromosome 8p12 was reported to be associated with PTC [27]. This association has been confirmed in multiple replication studies in Icelandic, Korean, Japanese, and Chinese populations (Table 5) [28, 63, 67]. The rs2439302 variant has been confirmed as a PTC risk variant with odds ratios ranging from 1.2 to 1.4. This variant has also been correlated with multifocality and lymph node metastasis in PTC patients [40]. Another variant of NRG1 locus (SNP rs2466076) was found to have an OR of 1.32 among 3001 NMTC cases and 287,550 controls [65]. Both SNPs are located in the intronic regions of the neuregulin 1 (NRG1) gene. The NRG1 gene encodes a membrane glycoprotein that mediates cell-cell signalling and plays a critical role in the growth and development of multiple organ systems.

Table 5 NRG1 variants associated with hereditary thyroid cancer

Additionally, NRG1 dysregulation is closely linked to PI3K-AKT and MAPK signalling pathways and has been demonstrated to be involved in tumourigenesis of both malignant and benign thyroid tumours [69, 70]. He et al. evaluated candidate functional variants of NRG1. The [G] risk allele (rs2439302) was associated with higher expression of the three tested isoforms in normal thyroid tissue. The authors proposed these isoforms as contributing factors to higher PTC risk through allele-specific enhancer-mediated transcriptional regulation of NRG1 [71]. NRG1 expression was also shown to be essential for PTC cell proliferation through protection from reactive oxygen species (ROS) damage by nuclear factor E2-related factor 2 (NRF2). Therefore, NRG1 can also be useful as a potential therapeutic target for PTC patients [72, 73]. Guibon et al. performed fine-mapping of the 8p12 (NRG1) locus in Europeans, Melanesians and Polynesians populations and identified rs2439304 associated with DTC (OR = 1.2). This variant had the highest posterior probability (PP) of causality in the three ethnic groups based on expression Quantitative Trait Locus (eQTL) data at this locus [74]. NRG1 variants show stronger association in the Korean population compared to the European populations, suggesting a potential Korean-specific marker for DTC [29, 75].

DIRC3

DIRC3 (Disrupted In Renal Carcinoma 3) is an RNA gene affiliated with the lncRNA class of RNAs. Several diseases have been associated with DIRC3, including renal cell, breast, and thyroid carcinoma. Multiple reports have demonstrated the prognostic significance of the rs966423 variant of the DIRC3 gene and its pathogenic effects in DTC cases. DIRC3 was first identified in 2003 as a fusion transcript involved in familial renal carcinoma. Although the function of DIRC3 is still unknown, it is thought to have tumour suppressor activity [76]. In a GWAS with 561 Icelandic individuals with thyroid cancer cases and 40,013 controls, DIRC3 variants were associated both with thyroid cancer risk and thyroid stimulating hormone levels. One variant that was significantly correlated with PTC was rs966423 with an OR of 1.34 [27]. In the replication studies by Köhler et al. (2013) and Son et al. (2017) three other intronic variants in this gene were identified in DTC and PTC cases with low penetrance (Table 6) [29, 30]. However, Mankickova et al. were unable to establish an association between rs966423 and thyroid cancer in a European population, suggesting inter-population heterogeneity in thyroid cancer susceptibility [33]. Patients homozygous for the T allele of rs966423 have a 6.4% higher mortality risk compared to CC/CT carriers (P = 0.017) [77]. Additionally, CT genotype carriers were associated with extrathyroidal extension and more advanced T stage [78]. On the contrary, a recent study in 1466 DTC patients reported no association between any genotype at the rs966423 SNP and overall mortality and response to therapy [79]. A recent GWAS analysis also identified five novel variants including rs11693806 as a non-coding variant located close to DIRC3 in a large sample of 2637 European ancestry cases and 134,811 European ancestry controls [28]. Another study by Guibon et al. identified rs16857609 as a novel variant located near DIRC3. This SNP was associated with DTC in the European population (OR = 1.4, p = 1.9 × 10− 10) [74]. Future studies should replicate the findings of the known DIRC3 variants and confirm their association with PTC pathogenesis.

Table 6 DIRC3 variants associated with hereditary thyroid cancer

Polygenic risk score

As reviewed in the previous section and summarized in Table 1, the genome-wide association studies identified many low penetrant risk alleles for thyroid cancer. Single genetic variants with such low-risk alleles do not explain the clustering of thyroid cancer in families. Consequently, polygenic risk scores (PRS) have been developed to consider panels of SNPs to calculate their additive risk for thyroid cancer. The integration of PRS with family history can potentially improve identifying people at risk for developing thyroid cancer in various populations. A recent study investigated the combined genetic effects of 10 well-established SNPs (rs12129938, rs11693806, rs6793295, rs73227498, rs2466076, rs1588635, rs7902587, rs368187, rs116909374, and rs2289261) associated with PTC by evaluating their PRS with data from previous GWAS from United States, Iceland, and the United Kingdom. Their results indicate a 6.9-fold greater risk for thyroid cancer for patients in the top decile of the ten common SNPs polygenic risk scores compared to the bottom decile [80]. Similarly, Hoang et al. investigated the value of PRS for thyroid cancer in a Korean population. In this study, a family history of thyroid cancer (OR = 2.96), obesity (OR = 1.72), weighted (OR = 1.56), and unweighted PRS (OR = 1.46) were associated with thyroid cancer susceptibility [81]. The PRS of 12 thyroid cancer-associated SNPs (rs11693806, rs2466076, rs1588635, rs368187, rs116909374, rs12129938, rs6793295, rs73227498, rs7902587, rs2289261, and rs56062135) was investigated in 2370 childhood cancer survivors with an European ancestry. Similar to previous findings, the hazard ratio for developing secondary thyroid cancer by one standard deviation increase in the PRS was 1.57 (95% CI = 1.25–1.83; P <  0.001) [82]. Likewise, in a phenome-wide association study of 472 thyroid cancer patients with European ancestry, a PRS of 9 SNPs exhibited a strong association with thyroid cancer (OR = 3.2) when the top PRS quartile was compared to the bottom quartile [83]. In a study by Wang et al., Individuals with African ancestry who were in the top PRS quintile of 5 SNPs had a 30% greater chance of thyroid cancer (OR = 1.3) than those in the lowest quintile [84]. Additionally, in a study with 495 thyroid cancer patients and 56,439 controls by Song et al., the PRS of 6 SNPs (rs6759952, rs13059137, rs7834206, rs72616195, rs1369535, rs11175834) increased thyroid cancer risk by a factor of 3.9 when comparing high PRS tertile with low PRS tertile [85]. Given the presented findings, PRS has the potential to identify individuals at a higher risk of thyroid cancer. However, studies with larger sample sizes and more inclusive PRS with wide varieties of SNPs are required for determining the optimal PRS model for thyroid cancer.

Rare germline mutations in families with Non-syndromic familial non-medullary thyroid cancer

Non-syndromic familial non-medullary thyroid cancer (NSFNMTC) accounts for 95% of FNMTC cases. The genetic risk factors of non-syndromic FNMTC are poorly understood compared to familial NMTC associated with hereditary syndromes (syndromic NMTC). In addition to FOXE1, HABP2, NRG1, SRGAP1, DIRC3, TITF1/ NKX2.1 and PTCSC3, multiple other genes and chromosomal loci have been linked to families affected by non-syndromic FNMTC in linkage studies and/or whole-exome/whole-genome sequencing studies. The identified mutations are present in only a subset of FNMTC kindreds and require further validation studies. Table 7 summarizes multiple studies that investigated the genetic component of FNMTC in families with NSFNMTC.

Table 7 Genes and chromosomal loci linked to non-syndromic familial non-medullary thyroid cancer

Syndromic familial non-medullary thyroid cancer

Hereditary syndromes (syndromic FNMTC) with various clinical features may be associated with approximately 5% of familial non-medullary thyroid cancer cases (Table 8). In addition to the implicated syndrome or disease symptoms, patients with syndromic FNMTC may develop cancers of non-thyroidal origin as well.. In a recent study, Zhou et al. checked twenty-five candidate NMTC susceptibility genes against six genetic resources including ClinGen, NCCN guidelines, OMIM, Genetics Home Reference, GeneCards, and Gene-NCBI. These susceptibility genes were assessed based on gene-disease association from previous studies. Subsequently, 12 genes (APC, DICER1, FOXE1, HABP2, NKX2–1, PRKAR1A, PTEN, SDHB, SDHD, SRGAP1, CHEK2, and SEC23B) were verified as NMTC susceptibility genes. Seventy-nine diseases were associated with these 12 susceptibility loci, some of which are causative genetic components of syndromic FNMTC, while others have been implicated in non-syndromic FNMTC [109]. The predominant syndromes that may lead to the development of syndromic NMTC are familial adenomatous polyposis (FAP), Cowden’s disease, Carney’s complex type 1, Werner’s syndrome, DICER1 syndrome, Li-Fraumeni syndrome, PTEN hamartoma tumour syndrome, Peutz-Jeghers syndrome, Bannayan-Riley-Ruvalcaba syndrome, Ataxia-telangiectasia, and Pendred syndrome. Syndromic FNMTC susceptibility genes and their highly penetrant mutations could be of great value for screening at-risk individuals, thereby making early diagnosis and selecting appropriate treatment possible. It is important for clinicians to recognize the phenotypes of these syndromes so that genetic counselling can be initiated to enable surveillance for associated malignancies and genetic testing of family members. Additionally, more frequent screening is warranted for first-degree family members of patients affected by syndromic FNMTC.

Table 8 Hereditary syndromes associated with thyroid cancers of follicular cell origin

Familial adenomatous polyposis (FAP) and Gardner’s syndrome

Familial adenomatous polyposis (FAP) is an autosomal dominant disease caused by loss-of-function mutations of the APC tumour suppressor gene located on chromosome 5q21. The classic type of FAP is characterized by the development of multiple benign polyps lining the mucosa of the gastrointestinal tract, particularly the colon. Untreated polyps can become malignant with an early age of onset. Papillary thyroid carcinomas are seen in some families affected by FAP [110, 111]. In fact, patients with FAP have a 160-fold greater risk of PTC compared to the general population. The prevalence of thyroid cancer among patients with FAP is 2.6%. These thyroid cancers have a unique cribriform pattern on histologic examination and occur more commonly at a young age (< 30 years) in women (95%) [112, 113].

More than 60% of APC pathogenic mutations have been identified in the mutation cluster region between codons 1284 and 1580 [114, 115]. Most female patients with FAP and PTC also have a RET somatic mutation in addition to APC germline mutations in their tumours [116]. The APC gene encodes a multidomain protein that plays a significant role in tumour suppression by negatively regulating the WNT signalling pathway. Loss of APC function results in inappropriate activation of this pathway which results in cancer progression [117].

Werner’s syndrome

Werner’s syndrome is an autosomal recessive disease characterized by premature aging, scleroderma-like skin changes, cataracts, subcutaneous calcifications, muscular atrophy, diabetes, and a high incidence of neoplasms, including thyroid neoplasms. Werner’s syndrome has been linked to mutations of the WRN gene on chromosome 8p11–21. This gene encodes a member of the RecQ subfamily of DNA helicase proteins that is important in maintaining genome stability by regulating DNA repair, replication, transcription, and telomere maintenance [118]. Thyroid cancer was observed in 16% of 189 patients with Werner syndrome in a Japanese case series. Follicular thyroid cancer was more common, followed by papillary and anaplastic thyroid cancers among these patients [119].

Carney complex

Carney complex (CNC) is an autosomal dominant disease caused by mutations in the PRKAR1 tumour suppressor gene mapped to chromosome 17q22–24 [120]. A loss of function mutation in PRKAR1A can lead to increased PKA signalling [121]. Additionally, this gene can fuse to the RET proto-oncogene by gene rearrangement and forming a thyroid tumour-specific chimeric oncogene known as PTC2. A loss of function mutation in PARKARI causes increased PKA signalling, leading to AMP-activated kinase (AMPK) activation through LKB1 kinase and increasing mTOR signalling [122]. As a result, patients may present with acromegaly, spotty skin pigmentation, an increased risk of cardiac and mucocutaneous myxomas, and a variety of tumours involving endocrine organs.

Additionally, about 60% of patients affected by CNC will develop thyroid tumours that range from follicular hyperplasia to multiple types of thyroid cancer, with follicular adenoma as the most common finding [7]. In a study by Stratakis et al. the prevalence of thyroid nodules and cancers in a series of 338 Carney’s complex patients was 5%, including follicular adenomas, PTC, follicular variant PTC (FvPTC), and FTC [123]. Patients affected by CNC should undergo surveillance using frequent ultrasound and biopsies to increase the likelihood of treatment success.

DICER1 syndrome

DICER1 syndrome, also known as pleuropulmonary blastoma syndrome and dysplasia syndrome, is an autosomal dominant genetic disorder that predisposes individuals to various conditions, including benign and malignant tumours of different origins. Germline mutations of the DICER1 gene located on 14q32.13 are detected in endocrine tumours (thyroid, parathyroid, pituitary, pineal gland, endocrine pancreas, paragangliomas, medullary, adrenocortical, ovarian, and testicular tumours).

The DICER1 gene is a member of the ribonuclease III (RNaseIII) family involved in the generation of microRNAs (miRNAs) and modulates gene expression by interfering with mRNA function. DICER1 germline loss-of-function mutations disrupt the correct timing and expression of miRNA production necessary for normal thyroid differentiation and function [124, 125]. Khan et al. investigated the risk of thyroid cancer in 145 individuals with DICER1 germline mutations and 135 family controls from 48 families. This group reported a 16-fold increased risk of thyroid cancer, with all the cases harbouring germline and somatic pathogenic DICER1 mutations [126].

Thyroid abnormalities are common in DICER1 syndrome with multinodular goiter seen frequently in many families with a germline DICER1 mutation. Thus, familial MNG is highly suggestive of DICER1 syndrome. In contrast, differentiated thyroid carcinoma (DTC) was infrequently seen in pedigrees with germline DICER1 mutation. However, multiple differentiated thyroid carcinomas have been found in three children with a history of prior chemotherapy and radiation exposure for the treatment of pleuropulmonary blastoma (PPB). As a result, there has been considerable speculation on a possible link between chemotherapeutic agents and an increased risk of differentiated thyroid cancer due to somatic DICER1 mutations [127]. More recently, a family study reported differentiated thyroid cancer and MNG in six individuals from a family with DICER1 pathogenic mutations and no history of chemotherapy [128].

PTEN hamartoma tumour syndromes

PTEN hamartoma tumour syndrome (PHTS) consists of a group of disorders caused by germline mutations in the phosphatase and tensin homolog (PTEN) gene located at 10q23.31. They include Cowden syndrome, Bannayan-Riley-Ruvalcaba syndrome (BRRS), PTEN-related Proteus syndrome, Proteus-like syndrome, and adult Lhermitte-Duclos disease (LDD) [129, 130]. Approximately 6 to 38% of PHTS patients develop thyroid cancer with a median age of diagnosis of 31–37 years, indicating a risk 51 to 72 times higher than those without PHTS [130]. Therefore, the presence of a PTEN mutation justifies surveillance with annual neck palpation and ultrasound imaging starting at age 10 [131, 132].

PTEN is a phosphatase that counteracts the phosphatidylinositol 3-kinase (PI3K)/AKT signalling pathways. The tumour-suppressor activity of PTEN is thought to be associated with lipid dephosphorylation at the plasma membrane. An inactivated PTEN gene may increase PIP3 levels leading to AKT activation and mTOR signalling which in turn upregulates cell proliferation and survival while decreasing apoptosis [133, 134].

Patients with Cowden syndrome may suffer from breast, endometrium, colon, thyroid, and kidney tumours in addition to NMTC due to PTEN mutations. At least two-thirds of patients with this syndrome are affected by thyroid disease, often before the age of 20. In addition, approximately 10% of patients with Cowden syndrome will develop thyroid cancer (FTC or PTC) in their lifetime [7], [135]. In a study of 664 patients with Cowden syndrome (CS) or Cowden-like syndrome (CLS), 55.1% of the thyroid cancer cases were of classical papillary subtype. In this cohort, 5.4% of the CS and CLS patients had PTEN germline mutations. About 4% of the patients that did not harbor PTEN mutations tested positive for SDHB-D mutations and 2.3% tested positive for KLLN promoter methylation [136]. The SDHB-D genes located on chromosome 1p36.13 encodes succinate dehydrogenase (SDH). Its germline variants can result in the upregulation of the AKT and MAPK pathways, similar to PTEN mutations that can drive tumour formation [137]. KLLN is a tumour suppressor gene located upstream of PTEN. KLLN promoter methylation downregulates its transcription and disrupts p53-mediated activation of KLLN [138]. In another study, targeted sequencing of 91 CS and CLS probands without PTEN, SDHB-D and KLLN mutations, revealed PIK3CA germline mutations in 8.8% and AKT1 germline mutations in 2.2% of cases [139]. PIK3CA is a gene located on chromosome 3q26.32 and encodes p110a, the catalytic subunit of PI3K, which adds a phosphate to phosphatidylinositol-4,5-biphosphate (PIP2) to form phosphatidylinositol-3,4,5-triphosphate (PIP3) at the cellular membrane. PIP3 recruits AKT1 to the cell membrane. Subsequently, activated AKT phosphorylates downstream protein effectors, including the mammalian target of rapamycin (mTOR), which has an established role in human cancers [140].

Whole exome sequencing of a CS proband with the other family members affected with thyroid cancer across 4 generations was performed in 2015. Although all individuals tested negative for PTEN, SDHB-D, KLLN hypermethylation, PIK3CA and AKT1, several novel candidate genes were identified. All family members with CS shared 3 genes with heterozygous missense variants, C16orf72 (c.253 T > C,p.Ser85Pro), PTPN2 (c.1204G > A,p.Ala402Thr) and SEC23B (c.1781 T > G, p.Val594Gly). All 3 genes were sequenced in 96 unrelated CS probands with thyroid cancer, and germline heterozygous SEC23B variant was detected in 3 probands (3.1%). SEC23B encodes Sec23 Homolog B, a component of coat protein complex II (COPII) responsible for transporting proteins from the endoplasmic reticulum (ER) to the Golgi apparatus [141].

Bannayan-Rubalcaba-Riley syndrome (BRRS) is an overgrowth disorder with germline PTEN tumour-suppressor gene involvement in 60% of cases. BRRS involves macrocephaly, pigmented maculae of the glans penis, and benign mesodermal hamartomas. About 30% of BRRS patients may have diseases of thyroid origin, including, NMTC, thyroid adenoma, MNG, and Hashimoto’s disease [136, 142].

Rare syndromes associated with NMTC

Studies involving patients with syndromic FNMTC have the power to add to the list of possible thyroid cancer susceptibility loci and help the identification of key players in thyroid tumorigenesis. Case reports and familial studies have identified multiple rare syndromes associated with a risk for NMTC. Li-Fraumeni syndrome caused by mutations in TP53 gene and presents with a high risk of cancers with bone, breast, adrenal gland, and nervous system origins, with a lifetime cancer risk of > 70% for men and > 90% for women [143]. Formiga et al. established the presence of thyroid cancer in 193 Li-Fraumeni Syndrome (LFS) patients. 101 Out of 193 LFS cases, 101 were carriers of the Brazilian TP53 p.R337H mutation with 10.9% of cases exhibiting papillary thyroid carcinoma tumours [144].

Pendred syndrome is an autosomal recessive disorder characterized by bilateral sensorineural deafness and goitre caused by mutations in the SLC26A4 (PDS) gene (7q12–34) [145]. The protein product of the SLC26A4 gene is pendrin, a surface anion channel found on the apical membrane of thyroid follicular cells. A loss of function mutation in SLC26A4 may disrupt iodine transport and result in goitre and hypothyroidism [146]. Additionally, follicular thyroid cancer, Hürthle cell adenoma, MNG, and fvPTC have been observed in families affected by Pendred syndrome [147,148,149,150]. Untreated congenital hypothyroidism, chronic stimulation by thyroid-stimulating hormone, and additional genetic alterations may also be involved in the formation of thyroid cancer in pendred patients [145, 149, 151].

Ataxia-telangiectasia syndrome is an autosomal recessive disorder caused by mutations in the (Ataxia Telangiectasia, Mutated) ATM gene on 11q22–23. The ATM gene encodes a member of the phosphatidylinositol 3-kinase family and plays a role in cellular responses to DNA breaks and oxidative stress. Patients with Ataxia-telangiectasia may present with cerebral ataxia, immunodeficiency, telangiectasia, radiation sensitivity, thymic atrophy, and various malignancies, particularly those with lymphoid origin [152]. Furthermore, mutations in the ATM gene have been implicated in PTC and fvPTC [153,154,155,156,157]. A Danish population-based study of 10,324 individuals identified an association between heterozygosity at ATM Ser707Pro and thyroid/endocrine cancer (HR = 10) [158]. Additionally, using whole-genome sequencing and genome-wide linkage analysis, Wang et al. identified ATM variants in 2 of 17 families affected by FNMTC [92].

Peutz-Jeghers syndrome (PJS) is an autosomal dominant disorder characterized by hamartomatous polyps, mucocutaneous hyperpigmentation, and a 4-fold increase in cancer risk compared to the general population [159]. Mutations in the STK11 (serine/threonine-protein kinase 11 alias LKB1) gene (19p13.3) have been implicated as a causative agent for PJS. The protein product of the STK11 gene is a serine-threonine kinase involved in second messenger signal transduction and AMPK inhibition [160]. Additionally, PJS has been associated with multiple cases of thyroid cancer of PTC, FTC, tall cell variant PTC, and fvPTC subtypes [159, 161, 162]. Papillary Renal Neoplasia (PRN) [101] and McCune–Albright syndrome [163] are two other rare disorders associated with risk of thyroid cancer. Nevertheless, only a few families have been affected by both thyroid cancer and the mentioned syndromes.

Conclusion

Non-medullary thyroid cancer originates from follicular cells of the thyroid gland and accounts for the majority of thyroid cancers. The genetic component of NMTC tumourigenesis is strong but poorly understood, especially for familial NMTC. This review aimed to summarize the current understanding of genetic predisposition to NMTC by looking at genetic variants implicated in familial and sporadic NMTC. Increasing evidence suggests that mutations in the FOXE1 gene have moderate to high penetrance. On the other hand, there is a lack of strong evidence for the role of HABP2 mutations. Thus, further research is needed to clarify its role as a susceptibility gene in NMTC. Most of the mutations in TITF1/ NKX2.1, PTCSC3, SRGAP1, NRG1, DIRC3 genes are low penetrant mutations. Although each low penetrant mutation does not seem to have clinical significance alone, a combination of these mutations could have clinical importance regarding hereditary NMTC.

Different germlines variants are only observed in small groups of FNMTC patients and may not be present in all affected family members within a kindred. Likewise, due to the lack of interventional screening programs, there are no genetic tests available to identify individuals at risk of FNMTC. As a result, the National Comprehensive Cancer Network (NCCN), the American Thyroid Association (ATA), and the European Society for Medical Oncology (ESMO) provide no recommendations on using genetic testing for screening at-risk family members of FNMTC patients [164,165,166]. We also do not think that we have enough evidences supporting the application of genetic screening for certain genes among patients with FNMTC, unless patients medical history and family history suggest a syndromic NMTC that should be tested for the related gene(s).

Further multi-center studies with larger cohorts and stricter inclusion criteria using targeted sequencing or whole exome/genome sequencing are needed to better understand the clustering pattern seen in the families with NMTC. Identification of NMTC susceptibility genes could potentially result in determining targeted treatment options for NMTC patients. Likewise, alternative hereditary mechanisms such as epigenetic modifications may also be involved in the pathogenesis of FNMTC and requires further research. Additionally, identifying new NMTC-associated genetic loci and research on the known implicated variants can improve our understanding of NMTC tumorigenesis in general, which could eventually result in earlier diagnosis and more effective treatment options for sporadic NMTC.

Availability of data and materials

Not applicable.

Abbreviations

NMTC:

Non-medullary thyroid cancer

DTC:

Differentiated thyroid cancer

PTC:

Papillary thyroid cancer

FNMTC:

Familial non-medullary thyroid cancer

GWAS:

Genome wide association studies

OR:

Odd Ratio

FOXE1 :

Forkhead box E1

SNP:

Single nucleotide polymorphism

TTF-1:

Thyroid-specific transcription factor-1

NKX2.1:

NK2 homeobox 1

lncRNA:

Long non-coding RNA

MNG:

Multinodular goiter

SRGAP1:

Slit-Robo GTPase-activating protein 1

HABP2 :

Hyaluronan-Binding Protein 2

PTCSC2:

PTC susceptibility candidate 2

PTCSC3:

PTC susceptibility candidate 3

PRS:

Polygenic risk score

AMPK:

AMP-activated kinase

DIRC3:

Disrupted In Renal Carcinoma 3

NRG1 :

Neuregulin 1

PP:

Posterior probability of causality

MYH9:

Myosin-9

FAP:

Familial adenomatous polyposis

FTC:

Follicular thyroid cancer

ATC:

Anaplastic thyroid cancer

CS:

Cowden syndrome

CLS:

Cowden-like syndrome

CNC:

Carney complex

SDH:

Succinate dehydrogenase

PIP2:

Phosphatidylinositol-4,5-biphosphate

PIP3:

Phosphatidylinositol-3,4,5-triphosphate

PPB:

Pleuropulmonary blastoma

PI3K:

Phosphatidylinositol 3-kinase

mTOR:

Mammalian target of rapamycin

ROS:

Reactive oxygen species

eQTL:

expression Quantitative Trait Locus

PHTS:

PTEN hamartoma tumour syndrome

BRRS:

Bannayan-Riley-Ruvalcaba syndrome

LDD:

Lhermitte-Duclos disease

COPII:

Coat protein complex II

ER:

Endoplasmic reticulum

RNaseIII:

Ribonuclease III

miRNAs:

MicroRNAs

LFS:

Li-Fraumeni Syndrome

References

  1. Chrisoulidou A, Boudina M, Tzemailas A, Doumala E, Iliadou PK, Patakiouta F, et al. Histological subtype is the most important determinant of survival in metastatic papillary thyroid cancer. Thyroid Res. 2011;4(1):12.

    PubMed  PubMed Central  Google Scholar 

  2. Wiltshire JJ, Drake TM, Uttley L, Balasubramanian SP. Systematic review of trends in the incidence rates of Thyroid Cancer. Thyroid. 2016;26(11):1541–52.

    PubMed  Google Scholar 

  3. Pal T, Vogl FD, Chappuis PO, Tsang R, Brierley J, Renard H, et al. Increased risk for nonmedullary thyroid cancer in the first degree relatives of prevalent cases of nonmedullary thyroid cancer: a hospital-based study. J Clin Endocrinol Metab. 2001;86(11):5307–12.

    CAS  PubMed  Google Scholar 

  4. American Cancer Society. Cancer Facts & Figures 2017. Atlanta; American Cancer Society; 2017.

  5. Xing M. Molecular pathogenesis and mechanisms of thyroid cancer. Nat Rev Cancer. 2013;13(3):184–99.

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Nosé V. Familial thyroid cancer: a review. Mod Pathol. 2011;24(Suppl 2):S19–33.

    PubMed  Google Scholar 

  7. Peiling Yang S, Ngeow J. Familial non-medullary thyroid cancer: unraveling the genetic maze. Endocr Relat Cancer. 2016;23(12):R577–95.

    PubMed  Google Scholar 

  8. Guilmette J, Nosé V. Hereditary and familial thyroid tumours. Histopathology. 2018;72(1):70–81. https://doi.org/10.1111/his.13373.

    Article  PubMed  Google Scholar 

  9. Hińcza K, Kowalik A, Kowalska A. Current Knowledge of Germline Genetic Risk Factors for the Development of Non-Medullary Thyroid Cancer. Genes (Basel). 2019;10(7):4882.

    Google Scholar 

  10. Czene K, Lichtenstein P, Hemminki K. Environmental and heritable causes of cancer among 9.6 million individuals in the Swedish family-Cancer database. Int J Cancer. 2002;99(2):260–6.

    CAS  PubMed  Google Scholar 

  11. Bonora E, Tallini G, Romeo G. Genetic predisposition to familial nonmedullary Thyroid Cancer: an update of molecular findings and state-of-the-art studies. J Oncol. 2010;2010:385206.

    PubMed  PubMed Central  Google Scholar 

  12. Goldgar DE, Easton DF, Cannon-Albright LA, Skolnick MH. Systematic population-based assessment of cancer risk in first-degree relatives of cancer probands. J Natl Cancer Inst. 1994;86(21):1600–8.

    CAS  PubMed  Google Scholar 

  13. Dong C, Hemminki K. Modification of cancer risks in offspring by sibling and parental cancers from 2,112,616 nuclear families. Int J Cancer. 2001;92(1):144–50.

    CAS  PubMed  Google Scholar 

  14. Hansen PS, Brix TH, Bennedbaek FN, Bonnema SJ, Kyvik KO, Hegedüs L. Genetic and environmental causes of individual differences in thyroid size: a study of healthy Danish twins. J Clin Endocrinol Metab. 2004;89(5):2071–7.

    CAS  PubMed  Google Scholar 

  15. Hansen PS, Brix TH, Bennedbaek FN, Bonnema SJ, Iachine I, Kyvik KO, et al. The relative importance of genetic and environmental factors in the aetiology of thyroid nodularity: a study of healthy Danish twins. Clin Endocrinol. 2005;62(3):380–6. https://doi.org/10.1111/j.1365-2265.2005.02230.x.

    Article  Google Scholar 

  16. Lin H-T, Liu F-C, Lin S-F, Kuo C-F, Chen Y-Y, Yu H-P. Familial Aggregation and Heritability of Nonmedullary Thyroid Cancer in an Asian Population: A Nationwide Cohort Study. J Clin Endocrinol Metab. 2020;105(7):dgaa191.

    PubMed  Google Scholar 

  17. Moses W, Weng J, Kebebew E. Prevalence, Clinicopathologic features, and somatic genetic mutation profile in familial versus sporadic nonmedullary Thyroid Cancer. Thyroid. 2011;21(4):367–71.

    PubMed  PubMed Central  Google Scholar 

  18. Mazeh H, Benavidez J, Poehls JL, Youngwirth L, Chen H, Sippel RS. In patients with thyroid cancer of follicular cell origin, a family history of nonmedullary thyroid cancer in one first-degree relative is associated with more aggressive disease. Thyroid. 2012;22(1):3–8.

    PubMed  Google Scholar 

  19. Capezzone M, Robenshtok E, Cantara S, Castagna MG. Familial non-medullary thyroid cancer: a critical review. J Endocrinol Investig. 2021;44(5):943–50.

    CAS  Google Scholar 

  20. Cao J, Chen C, Chen C, Wang Q-L, Ge M-H. Clinicopathological features and prognosis of familial papillary thyroid carcinoma--a large-scale, matched, case-control study. Clin Endocrinol. 2016;84(4):598–606.

    Google Scholar 

  21. Maxwell EL, Hall FT, Freeman JL. Familial non-medullary thyroid cancer: a matched-case control study. Laryngoscope. 2004;114(12):2182–6.

    PubMed  Google Scholar 

  22. Pitoia F, Cross G, Salvai ME, Abelleira E, Niepomniszcze H. Patients with familial non-medullary thyroid cancer have an outcome similar to that of patients with sporadic papillary thyroid tumors. Arq Bras Endocrinol Metabol. 2011;55(3):219–23.

    PubMed  Google Scholar 

  23. Capezzone M, Fralassi N, Secchi C, Cantara S, Brilli L, Pilli T, et al. Long-term clinical outcome in familial and sporadic papillary Thyroid carcinoma. Eur Thyroid J. 2020;9(4):213–20.

    PubMed  PubMed Central  Google Scholar 

  24. Park YJ, Ahn HY, Choi HS, Kim KW, Park DJ, Cho BY. The long-term outcomes of the second generation of familial nonmedullary thyroid carcinoma are more aggressive than sporadic cases. Thyroid. 2012;22(4):356–62.

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Klubo-Gwiezdzinska J, Yang L, Merkel R, Patel D, Nilubol N, Merino MJ, et al. Results of screening in familial Non-medullary Thyroid Cancer. Thyroid. 2017;27(8):1017–24.

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Risch N. The genetic epidemiology of cancer: interpreting family and twin studies and their implications for molecular genetic approaches. Cancer Epidemiol Biomark Prev. 2001;10(7):733–41.

    CAS  Google Scholar 

  27. Gudmundsson J, Sulem P, Gudbjartsson DF, Jonasson JG, Masson G, He H, et al. Discovery of common variants associated with low TSH levels and thyroid cancer risk. Nat Genet. 2012;44(3):319–22.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Gudmundsson J, Thorleifsson G, Sigurdsson JK, Stefansdottir L, Jonasson JG, Gudjonsson SA, et al. A genome-wide association study yields five novel thyroid cancer risk loci. Nat Commun. 2017;8:14517.

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Son H-Y, Hwangbo Y, Yoo S-K, Im S-W, Yang SD, Kwak S-J, et al. Genome-wide association and expression quantitative trait loci studies identify multiple susceptibility loci for thyroid cancer. Nat Commun. 2017;8:15966.

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Köhler A, Chen B, Gemignani F, Elisei R, Romei C, Figlioli G, et al. Genome-wide association study on differentiated thyroid cancer. J Clin Endocrinol Metab. 2013;98(10):E1674–81.

    PubMed  Google Scholar 

  31. Figlioli G, Köhler A, Chen B, Elisei R, Romei C, Cipollini M, et al. Novel genome-wide association study-based candidate loci for differentiated thyroid cancer risk. J Clin Endocrinol Metab. 2014;99(10):E2084–92. https://doi.org/10.1210/jc.2014-1734.

    Article  CAS  PubMed  Google Scholar 

  32. Figlioli G, Chen B, Elisei R, Romei C, Campo C, Cipollini M, et al. Novel genetic variants in differentiated thyroid cancer and assessment of the cumulative risk. Sci Rep. 2015;5:8922.

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Mancikova V, Cruz R, Inglada-Pérez L, Fernández-Rozadilla C, Landa I, Cameselle-Teijeiro J, et al. Thyroid cancer GWAS identifies 10q26.12 and 6q14.1 as novel susceptibility loci and reveals genetic heterogeneity among populations. Int J Cancer. 2015;137(8):1870–8.

    CAS  PubMed  Google Scholar 

  34. Zuk O, Hechter E, Sunyaev SR, Lander ES. The mystery of missing heritability: genetic interactions create phantom heritability. Proc Natl Acad Sci U S A. 2012;109(4):1193–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Gudmundsson J, Sulem P, Gudbjartsson DF, Jonasson JG, Sigurdsson A, Bergthorsson JT, et al. Common variants on 9q22.33 and 14q13.3 predispose to thyroid cancer in European populations. Nat Genet. 2009;41(4):460–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Landa I, Ruiz-Llorente S, Montero-Conde C, Inglada-Pérez L, Schiavi F, Leskelä S, et al. The variant rs1867277 in FOXE1 gene confers thyroid cancer susceptibility through the recruitment of USF1/USF2 transcription factors. PLoS Genet. 2009;5(9):e1000637.

    PubMed  PubMed Central  Google Scholar 

  37. Tomaz RA, Sousa I, Silva JG, Santos C, Teixeira MR, Leite V, et al. FOXE1 polymorphisms are associated with familial and sporadic nonmedullary thyroid cancer susceptibility. Clin Endocrinol. 2012;77(6):926–33.

    CAS  Google Scholar 

  38. Pereira JS, da Silva JG, Tomaz RA, Pinto AE, Bugalho MJ, Leite V, et al. Identification of a novel germline FOXE1 variant in patients with familial non-medullary thyroid carcinoma (FNMTC). Endocrine. 2015;49(1):204–14.

    CAS  PubMed  Google Scholar 

  39. Bonora E, Rizzato C, Diquigiovanni C, Oudot-Mellakh T, Campa D, Vargiolu M, et al. The FOXE1 locus is a major genetic determinant for familial nonmedullary thyroid carcinoma. Int J Cancer. 2014;134(9):2098–107.

    CAS  PubMed  Google Scholar 

  40. Jendrzejewski J, Liyanarachchi S, Nagy R, Senter L, Wakely PE, Thomas A, et al. Papillary Thyroid carcinoma: association between germline DNA variant markers and clinical parameters. Thyroid. 2016;26(9):1276–84. https://doi.org/10.1089/thy.2015.0665.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. He H, Li W, Liyanarachchi S, Srinivas M, Wang Y, Akagi K, et al. Multiple functional variants in long-range enhancer elements contribute to the risk of SNP rs965513 in thyroid cancer. Proc Natl Acad Sci U S A. 2015;112(19):6128–33.

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Wang Y, He H, Li W, Phay J, Shen R, Yu L, et al. MYH9 binds to lncRNA gene PTCSC2 and regulates FOXE1 in the 9q22 thyroid cancer risk locus. Proc Natl Acad Sci U S A. 2017;114(3):474–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Gara SK, Jia L, Merino MJ, Agarwal SK, Zhang L, Cam M, et al. Germline HABP2 mutation causing familial nonmedullary Thyroid Cancer. N Engl J Med. 2015;373(5):448–55.

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Zhou EY, Lin Z, Yang Y. HABP2 mutation and nonmedullary Thyroid Cancer. N Engl J Med. 2015;373(21):2084–5.

    PubMed  Google Scholar 

  45. Sponziello M, Durante C, Filetti S. HABP2 mutation and nonmedullary Thyroid Cancer. N Engl J Med. 2015;373(21):2085–6.

    PubMed  Google Scholar 

  46. Tomsic J, He H, de la Chapelle A. HABP2 mutation and nonmedullary Thyroid Cancer. N Engl J Med. 2015;373(21):2086.

    PubMed  Google Scholar 

  47. Zhang T, Xing M. HABP2 G534E Mutation in Familial Nonmedullary Thyroid Cancer. J Natl Cancer Inst. 2016;108(6):djv415.

    PubMed  PubMed Central  Google Scholar 

  48. Sahasrabudhe R, Stultz J, Williamson J, Lott P, Estrada A, Bohorquez M, et al. The HABP2 G534E variant is an unlikely cause of familial non-medullary thyroid cancer. J Clin Endocrinol Metab. 2016;10(3):1098–103.

    PubMed  Google Scholar 

  49. de Randamie R, Martos-Moreno GÁ, Lumbreras C, Chueca M, Donnay S, Luque M, et al. Frequent and rare HABP2 variants are not associated with increased Susceptibility to familial nonmedullary Thyroid carcinoma in the Spanish population. Horm Res Paediatr. 2018;89(6):397–407.

    PubMed  Google Scholar 

  50. Kowalik A, Gąsior-Perczak D, Gromek M, Siołek M, Walczyk A, Pałyga I, et al. The p.G534E variant of HABP2 is not associated with sporadic papillary thyroid carcinoma in a polish population. Oncotarget. 2017;8(35):58304–8.

    PubMed  PubMed Central  Google Scholar 

  51. Cantara S, Marzocchi C, Castagna MG, Pacini F. HABP2 G534E variation in familial non-medullary thyroid cancer: an Italian series. J Endocrinol Investig. 2017;40(5):557–60.

    CAS  Google Scholar 

  52. de Mello LEB, Araujo AN, Alves CX, de Paiva FJP, Brandão-Neto J, Cerutti JM. The G534E variant in HABP2 is not associated with increased risk of familial nonmedullary thyroid cancer in Brazilian Kindreds. Clin Endocrinol. 2017;87(1):113–4.

    Google Scholar 

  53. Colombo C, Muzza M, Proverbio MC, Ercoli G, Perrino M, Cirello V, et al. Segregation and expression analyses of hyaluronan-binding protein 2 (HABP2): insights from a large series of familial non-medullary thyroid cancers and literature review. Clin Endocrinol. 2017;86(6):837–44.

    CAS  Google Scholar 

  54. Weeks AL, Wilson SG, Ward L, Goldblatt J, Hui J, Walsh JP. HABP2 germline variants are uncommon in familial nonmedullary thyroid cancer. BMC Med Genet. 2016;17(1):60.

    PubMed  PubMed Central  Google Scholar 

  55. Alzahrani AS, Murugan AK, Qasem E, Al-Hindi H. HABP2 gene mutations do not cause familial or sporadic Non-medullary Thyroid Cancer in a highly inbred middle eastern population. Thyroid. 2016;26(5):667–71. https://doi.org/10.1089/thy.2015.0537.

    Article  CAS  PubMed  Google Scholar 

  56. Shen C-T, Zhang G-Q, Qiu Z-L, Song H-J, Sun Z-K, Luo Q-Y. Targeted next-generation sequencing in papillary thyroid carcinoma patients looking for germline variants predisposing to the disease. Endocrine. 2019;64(3):622–31.

    CAS  PubMed  Google Scholar 

  57. Bohórquez ME, Estrada AP, Stultz J, Sahasrabudhe R, Williamson J, Lott P, et al. The HABP2 G534E polymorphism does not increase nonmedullary thyroid cancer risk in Hispanics. Endocr Connect. 2016;5(3):123–7.

    PubMed  PubMed Central  Google Scholar 

  58. Kimura S. Thyroid-specific transcription factors and their roles in thyroid cancer. J Thyroid Res. 2011;2011:710213.

    PubMed  PubMed Central  Google Scholar 

  59. Zhang X, Gu Y, Li Y, Cui H, Liu X, Sun H, et al. Association of rs944289, rs965513, and rs1443434 in TITF1/TITF2 with Risks of Papillary Thyroid Carcinoma and with Nodular Goiter in Northern Chinese Han Populations. Int J Endocrinol [Internet]. 2020 11 [cited 2021 Jun 7];2020. Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7036112/

  60. Ngan ESW, Lang BHH, Liu T, Shum CKY, So M-T, Lau DKC, et al. A germline mutation (A339V) in thyroid transcription factor-1 (TITF-1/NKX2.1) in patients with multinodular goiter and papillary thyroid carcinoma. J Natl Cancer Inst. 2009;101(3):162–75.

    CAS  PubMed  Google Scholar 

  61. Cantara S, Capuano S, Formichi C, Pisu M, Capezzone M, Pacini F. Lack of germline A339V mutation in thyroid transcription factor-1 (TITF-1/NKX2.1) gene in familial papillary thyroid cancer. Thyroid Res. 2010;3(1):4.

    PubMed  PubMed Central  Google Scholar 

  62. Jendrzejewski J, He H, Radomska HS, Li W, Tomsic J, Liyanarachchi S, et al. The polymorphism rs944289 predisposes to papillary thyroid carcinoma through a large intergenic noncoding RNA gene of tumor suppressor type. Proc Natl Acad Sci U S A. 2012;109(22):8646–51.

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Rogounovitch TI, Bychkov A, Takahashi M, Mitsutake N, Nakashima M, Nikitski AV, et al. The common genetic variant rs944289 on chromosome 14q13.3 associates with risk of both malignant and benign thyroid tumors in the Japanese population. Thyroid. 2015;25(3):333–40.

    CAS  PubMed  Google Scholar 

  64. He H, Bronisz A, Liyanarachchi S, Nagy R, Li W, Huang Y, et al. SRGAP1 is a candidate gene for papillary thyroid carcinoma susceptibility. J Clin Endocrinol Metab. 2013;98(5):E973–80.

    PubMed  PubMed Central  Google Scholar 

  65. Vega FM, Ridley AJ. Rho GTPases in cancer cell biology. FEBS Lett. 2008;582(14):2093–101.

    CAS  PubMed  Google Scholar 

  66. Etienne-Manneville S. Cdc42--the Centre of polarity. J Cell Sci. 2004;117(Pt 8):1291–300. https://doi.org/10.1242/jcs.01115.

    Article  CAS  PubMed  Google Scholar 

  67. Wang Y-L, Feng S-H, Guo S-C, Wei W-J, Li D-S, Wang Y, et al. Confirmation of papillary thyroid cancer susceptibility loci identified by genome-wide association studies of chromosomes 14q13, 9q22, 2q35 and 8p12 in a Chinese population. J Med Genet. 2013;50(10):689–95.

    CAS  PubMed  Google Scholar 

  68. Matsuse M, Takahashi M, Mitsutake N, Nishihara E, Hirokawa M, Kawaguchi T, et al. The FOXE1 and NKX2-1 loci are associated with susceptibility to papillary thyroid carcinoma in the Japanese population. J Med Genet. 2011;48(9):645–8.

    CAS  PubMed  Google Scholar 

  69. Talmage DA. Mechanisms of neuregulin action. Novartis Found Symp. 2008;289:74–84 discussion 84-93.

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Liyanarachchi S, Wojcicka A, Li W, Czetwertynska M, Stachlewska E, Nagy R, et al. Cumulative risk impact of five genetic variants associated with papillary thyroid carcinoma. Thyroid. 2013;23(12):1532–40.

    CAS  PubMed  PubMed Central  Google Scholar 

  71. He H, Li W, Liyanarachchi S, Wang Y, Yu L, Genutis LK, et al. The role of NRG1 in the predisposition to papillary Thyroid carcinoma. J Clin Endocrinol Metab. 2017;103(4):1369–79.

    PubMed Central  Google Scholar 

  72. Zhang T-T, Qu N, Sun G-H, Zhang L, Wang Y-J, Mu X-M, et al. NRG1 regulates redox homeostasis via NRF2 in papillary thyroid cancer. Int J Oncol. 2018;53(2):685–93.

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Laskin J, Liu SV, Tolba K, Heining C, Schlenk RF, Cheema P, et al. NRG1 fusion-driven tumors: biology, detection, and the therapeutic role of afatinib and other ErbB-targeting agents. Ann Oncol. 2020;31(12):1693–703.

    CAS  PubMed  Google Scholar 

  74. Guibon J, Sugier P-E, Kulkarni O, Karimi M, Bacq-Daian D, Besse C, et al. Fine–mapping of two differentiated thyroid carcinoma susceptibility loci at 2q35 and 8p12 in Europeans, Melanesians and Polynesians. Oncotarget. 2021;12(5):493–506.

    PubMed  PubMed Central  Google Scholar 

  75. Saenko VA, Rogounovitch TI. Genetic polymorphism predisposing to differentiated Thyroid Cancer: a review of major findings of the genome-wide association studies. Endocrinol Metab (Seoul). 2018;33(2):164–74.

    CAS  Google Scholar 

  76. Bodmer D, Schepens M, Eleveld MJ, Schoenmakers EFPM. Geurts van Kessel a. disruption of a novel gene, DIRC3, and expression of DIRC3-HSPBAP1 fusion transcripts in a case of familial renal cell cancer and t (2;3)(q35;q21). Genes Chromosomes Cancer. 2003;38(2):107–16.

    CAS  PubMed  Google Scholar 

  77. Świerniak M, Wójcicka A, Czetwertyńska M, Długosińska J, Stachlewska E, Gierlikowski W, et al. Association between GWAS-derived rs966423 genetic variant and overall mortality in patients with differentiated Thyroid Cancer. Clin Cancer Res. 2016;22(5):1111–9.

    PubMed  Google Scholar 

  78. Wei W-J, Lu Z-W, Wang Y, Zhu Y-X, Wang Y-L, Ji Q-H. Clinical significance of papillary thyroid cancer risk loci identified by genome-wide association studies. Cancer Genet. 2015;208(3):68–75. https://doi.org/10.1016/j.cancergen.2015.01.004.

    Article  CAS  PubMed  Google Scholar 

  79. Hińcza K, Kowalik A, Pałyga I, Walczyk A, Gąsior-Perczak D, Mikina E, et al. Does the TT Variant of the rs966423 Polymorphism in DIRC3 Affect the Stage and Clinical Course of Papillary Thyroid Cancer? Cancers (Basel) [Internet]. 2020 12 [cited 2021 Jun 8];12(2). Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7072153/

  80. Liyanarachchi S, Gudmundsson J, Ferkingstad E, He H, Jonasson JG, Tragante V, et al. Assessing thyroid cancer risk using polygenic risk scores. Proc Natl Acad Sci U S A. 2020;117(11):5997–6002.

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Hoang T, Nguyen Ngoc Q, Lee J, Lee EK, Hwangbo Y, Kim J. Evaluation of modifiable factors and polygenic risk score in thyroid cancer. Endocr Relat Cancer. 2021;28(7):481–94.

    PubMed  Google Scholar 

  82. Song N, Liu Q, Wilson CL, Sapkota Y, Ehrhardt MJ, Gibson TM, et al. Polygenic risk score improves risk stratification and prediction of subsequent Thyroid Cancer after childhood Cancer. Cancer Epidemiol Biomark Prev. 2021;30(11):2096–104.

    CAS  Google Scholar 

  83. Fritsche LG, Gruber SB, Wu Z, Schmidt EM, Zawistowski M, Moser SE, et al. Association of Polygenic Risk Scores for multiple cancers in a phenome-wide study: results from the Michigan genomics initiative. Am J Hum Genet. 2018;102(6):1048–61.

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Wang L, Desai H, Verma SS, Le A, Hausler R, Verma A, et al. Performance of polygenic risk scores for cancer prediction in a racially diverse academic biobank. Genet Med. 2021;S1098–3600(21):05367–3.

    Google Scholar 

  85. Song SS, Huang S, Park S. Association of Polygenetic Risk Scores Related to cell differentiation and inflammation with Thyroid Cancer risk and genetic interaction with dietary intake. Cancers (Basel). 2021;13(7):1510.

    CAS  Google Scholar 

  86. Ye F, Gao H, Xiao L, Zuo Z, Liu Y, Zhao Q, et al. Whole exome and target sequencing identifies MAP 2K5 as novel susceptibility gene for familial non-medullary thyroid carcinoma. Int J Cancer. 2019;144(6):1321–30.

    CAS  PubMed  Google Scholar 

  87. Bakhsh AD, Ladas I, Hamshere ML, Bullock M, Kirov G, Zhang L, et al. An InDel in phospholipase-C-B-1 is linked with Euthyroid multinodular goiter. Thyroid. 2018;28(7):891–901.

    CAS  PubMed  Google Scholar 

  88. Pasquali D, Torella A, Accardo G, Esposito D, Del Vecchio BF, Salvatore D, et al. BROX haploinsufficiency in familial nonmedullary thyroid cancer. J Endocrinol Investig. 2021;44(1):165–71.

    CAS  Google Scholar 

  89. Wilson TL-S, Hattangady N, Lerario AM, Williams C, Koeppe E, Quinonez S, et al. A new POT1 germline mutation-expanding the spectrum of POT1-associated cancers. Familial Cancer. 2017;16(4):561–6.

    CAS  PubMed  Google Scholar 

  90. Richard MA, Lupo PJ, Morton LM, Yasui YA, Sapkota YA, Arnold MA, et al. Genetic variation in POT1 and risk of thyroid subsequent malignant neoplasm: a report from the childhood Cancer survivor study. PLoS One. 2020;15(2):e0228887.

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Srivastava A, Miao B, Skopelitou D, Kumar V, Kumar A, Paramasivam N, et al. A Germline Mutation in the POT1 Gene Is a Candidate for Familial Non-Medullary Thyroid Cancer. Cancers (Basel). 2020;12(6):1441.

    CAS  Google Scholar 

  92. Wang Y, Liyanarachchi S, Miller KE, Nieminen TT, Comiskey DF, Li W, et al. Identification of rare variants predisposing to Thyroid Cancer. Thyroid. 2019;29(7):946–55.

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Orois A, Gara SK, Mora M, Halperin I, Martínez S, Alfayate R, et al. NOP53 as A Candidate Modifier Locus for Familial Non-Medullary Thyroid Cancer. Genes (Basel) [Internet]. 2019 7 [cited 2021 Jun 9];10(11). Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6896177/

  94. Máximo V, Botelho T, Capela J, Soares P, Lima J, Taveira A, 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 (Hürthle cell) tumours of the thyroid. Br J Cancer. 2005;92(10):1892–8.

    PubMed  PubMed Central  Google Scholar 

  95. Bonora E, Evangelisti C, Bonichon F, Tallini G, Romeo G. Novel germline variants identified in the inner mitochondrial membrane transporter TIMM44 and their role in predisposition to oncocytic thyroid carcinomas. Br J Cancer. 2006;95(11):1529–36.

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Tomsic J, He H, Akagi K, Liyanarachchi S, Pan Q, Bertani B, et al. A germline mutation in SRRM2, a splicing factor gene, is implicated in papillary thyroid carcinoma predisposition. Sci Rep. 2015;5:10566.

    PubMed  PubMed Central  Google Scholar 

  97. Sarquis M, Moraes DC, Bastos-Rodrigues L, Azevedo PG, Ramos AV, Reis FV, et al. Germline mutations in familial papillary Thyroid Cancer. Endocr Pathol. 2020;31(1):14–20.

    CAS  PubMed  Google Scholar 

  98. He H, Nagy R, Liyanarachchi S, Jiao H, Li W, Suster S, et al. A susceptibility locus for papillary thyroid carcinoma on chromosome 8q24. Cancer Res. 2009;69(2):625–31.

    CAS  PubMed  PubMed Central  Google Scholar 

  99. He H, Li W, Wu D, Nagy R, Liyanarachchi S, Akagi K, et al. Ultra-rare mutation in long-range enhancer predisposes to thyroid carcinoma with high penetrance. PLoS One. 2013;8(5):e61920.

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Suh I, Filetti S, Vriens MR, Guerrero MA, Tumino S, Wong M, et al. Distinct loci on chromosome 1q21 and 6q22 predispose to familial nonmedullary thyroid cancer: a SNP array-based linkage analysis of 38 families. Surgery. 2009;146(6):1073–80.

    PubMed  Google Scholar 

  101. Malchoff CD, Sarfarazi M, Tendler B, Forouhar F, Whalen G, Joshi V, et al. Papillary thyroid carcinoma associated with papillary renal neoplasia: genetic linkage analysis of a distinct heritable tumor syndrome. J Clin Endocrinol Metab. 2000;85(5):1758–64.

    CAS  PubMed  Google Scholar 

  102. Bignell GR, Canzian F, Shayeghi M, Stark M, Shugart YY, Biggs P, et al. Familial nontoxic multinodular thyroid goiter locus maps to chromosome 14q but does not account for familial nonmedullary thyroid cancer. Am J Hum Genet. 1997;61(5):1123–30.

    CAS  PubMed  PubMed Central  Google Scholar 

  103. McKay JD, Lesueur F, Jonard L, Pastore A, Williamson J, Hoffman L, et al. Localization of a susceptibility gene for familial nonmedullary thyroid carcinoma to chromosome 2q21. Am J Hum Genet. 2001;69(2):440–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  104. McKay JD, Thompson D, Lesueur F, Stankov K, Pastore A, Watfah C, et al. Evidence for interaction between the TCO and NMTC1 loci in familial non-medullary thyroid cancer. J Med Genet. 2004;41(6):407–12.

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Prazeres HJ, Rodrigues F, Soares P, Naidenov P, Figueiredo P, Campos B, et al. Loss of heterozygosity at 19p13.2 and 2q21 in tumours from familial clusters of non-medullary thyroid carcinoma. Familial Cancer. 2008;7(2):141–9.

    PubMed  Google Scholar 

  106. Cavaco BM, Batista PF, Sobrinho LG, Leite V. Mapping a new familial thyroid epithelial neoplasia susceptibility locus to chromosome 8p23.1-p22 by high-density single-nucleotide polymorphism genome-wide linkage analysis. J Clin Endocrinol Metab. 2008;93(11):4426–30.

    CAS  PubMed  Google Scholar 

  107. Canzian F, Amati P, Harach HR, Kraimps J-L, Lesueur F, Barbier J, et al. A gene predisposing to familial Thyroid tumors with cell Oxyphilia maps to chromosome 19p13.2. Am J Hum Genet. 1998;63(6):1743–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  108. Bevan S, Pal T, Greenberg CR, Green H, Wixey J, Bignell G, et al. A comprehensive analysis of MNG1, TCO1, fPTC, PTEN, TSHR, and TRKA in familial nonmedullary Thyroid Cancer: confirmation of linkage to TCO1. J Clin Endocrinol Metabol. 2001;86(8):3701–4.

    CAS  Google Scholar 

  109. Zhou J, Singh P, Yin K, Wang J, Bao Y, Wu M, et al. Non-medullary Thyroid Cancer Susceptibility Genes: Evidence and Disease Spectrum. Ann Surg Oncol [Internet]. 2021 3 [cited 2021 Jun 7]; Available from: https://doi.org/10.1245/s10434-021-09745-x.

  110. Soravia C, Sugg SL, Berk T, Mitri A, Cheng H, Gallinger S, et al. Familial adenomatous polyposis-associated Thyroid Cancer: a clinical, pathological, and molecular genetics study. Am J Pathol. 1999;154(1):127–35.

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Vriens MR, Suh I, Moses W, Kebebew E. Clinical features and genetic predisposition to hereditary nonmedullary thyroid cancer. Thyroid. 2009;19(12):1343–9.

    CAS  PubMed  Google Scholar 

  112. Fenton PA, Clarke SE, Owen W, Hibbert J, Hodgson SV. Cribriform variant papillary thyroid cancer: a characteristic of familial adenomatous polyposis. Thyroid. 2001;11(2):193–7.

    CAS  PubMed  Google Scholar 

  113. Chenbhanich J, Atsawarungruangkit A, Korpaisarn S, Phupitakphol T, Osataphan S, Phowthongkum P. Prevalence of thyroid diseases in familial adenomatous polyposis: a systematic review and meta-analysis. Familial Cancer. 2019;18(1):53–62.

    CAS  Google Scholar 

  114. Bertario L, Russo A, Sala P, Varesco L, Giarola M, Mondini P, et al. Multiple approach to the exploration of genotype-phenotype correlations in familial adenomatous polyposis. JCO. 2003;21(9):1698–707.

    CAS  Google Scholar 

  115. De Rosa M, Scarano MI, Panariello L, Morelli G, Riegler G, Rossi GB, et al. The mutation spectrum of the APC gene in FAP patients from southern Italy: detection of known and four novel mutations. Hum Mutat. 2003;21(6):655–6.

    PubMed  Google Scholar 

  116. Cetta F, Chiappetta G, Melillo RM, Petracci M, Montalto G, Santoro M, et al. The ret/ptc1 oncogene is activated in familial adenomatous polyposis-associated thyroid papillary carcinomas. J Clin Endocrinol Metab. 1998;83(3):1003–6.

    CAS  PubMed  Google Scholar 

  117. Cetta F, Montalto G, Gori M, Curia MC, Cama A, Olschwang S. Germline mutations of the APC gene in patients with familial adenomatous polyposis-associated thyroid carcinoma: results from a European cooperative study. J Clin Endocrinol Metab. 2000;85(1):286–92.

    CAS  PubMed  Google Scholar 

  118. Lauper JM, Krause A, Vaughan TL, Monnat RJ. Spectrum and risk of neoplasia in Werner syndrome: a systematic review. PLoS One. 2013;8(4):e59709.

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Ishikawa Y, Sugano H, Matsumoto T, Furuichi Y, Miller RW, Goto M. Unusual features of thyroid carcinomas in Japanese patients with Werner syndrome and possible genotype-phenotype relations to cell type and race. Cancer. 1999;85(6):1345–52.

    CAS  PubMed  Google Scholar 

  120. Sandrini F, Matyakhina L, Sarlis NJ, Kirschner LS, Farmakidis C, Gimm O, et al. Regulatory subunit type I-α of protein kinase a (PRKAR1A): a tumor-suppressor gene for sporadic thyroid cancer. Genes Chromosom Cancer. 2002;35(2):182–92.

    CAS  PubMed  Google Scholar 

  121. Kirschner LS, Sandrini F, Monbo J, Lin JP, Carney JA, Stratakis CA. Genetic heterogeneity and spectrum of mutations of the PRKAR1A gene in patients with the carney complex. Hum Mol Genet. 2000 Dec 12;9(20):3037–46.

    CAS  PubMed  Google Scholar 

  122. Kari S, Vasko VV, Priya S, Kirschner LS. PKA Activates AMPK Through LKB1 Signaling in Follicular Thyroid Cancer. Front Endocrinol (Lausanne). 2019;10:769.

    Google Scholar 

  123. Stratakis CA, Kirschner LS, Carney JA. Clinical and molecular features of the Carney complex: diagnostic criteria and recommendations for patient evaluation. J Clin Endocrinol Metab. 2001;86(9):4041–6.

    CAS  PubMed  Google Scholar 

  124. Slade I, Bacchelli C, Davies H, Murray A, Abbaszadeh F, Hanks S, et al. DICER1 syndrome: clarifying the diagnosis, clinical features and management implications of a pleiotropic tumour predisposition syndrome. J Med Genet. 2011;48(4):273–8.

    CAS  PubMed  Google Scholar 

  125. Fuziwara CS, Kimura ET. MicroRNAs in thyroid development, function and tumorigenesis. Mol Cell Endocrinol. 2017;456:44–50.

    CAS  PubMed  Google Scholar 

  126. Khan NE, Bauer AJ, Schultz KAP, Doros L, Decastro RM, Ling A, et al. Quantification of Thyroid Cancer and multinodular goiter risk in the DICER1 syndrome: a family-based cohort study. J Clin Endocrinol Metab. 2017;102(5):1614–22.

    PubMed  PubMed Central  Google Scholar 

  127. de Kock L, Sabbaghian N, Soglio DB-D, Guillerman RP, Park B-K, Chami R, et al. Exploring the association between DICER1 mutations and differentiated thyroid carcinoma. J Clin Endocrinol Metab. 2014;99(6):E1072–7.

    PubMed  Google Scholar 

  128. Rutter MM, Jha P, Schultz KAP, Sheil A, Harris AK, Bauer AJ, et al. DICER1 mutations and differentiated Thyroid carcinoma: Evidence of a direct association. J Clin Endocrinol Metab. 2016;101(1):1–5.

    CAS  PubMed  Google Scholar 

  129. Pilarski R, Burt R, Kohlman W, Pho L, Shannon KM, Swisher E. Cowden syndrome and the PTEN hamartoma tumor syndrome: systematic review and revised diagnostic criteria. J Natl Cancer Inst. 2013;105(21):1607–16.

    CAS  Google Scholar 

  130. Hendricks LAJ, Hoogerbrugge N, Schuurs-Hoeijmakers JHM, Vos JR. A review on age-related cancer risks in PTEN hamartoma tumor syndrome. Clin Genet. 2021;99(2):219–25.

    CAS  PubMed  Google Scholar 

  131. Bubien V, Bonnet F, Brouste V, Hoppe S, Barouk-Simonet E, David A, et al. High cumulative risks of cancer in patients with PTEN hamartoma tumour syndrome. J Med Genet. 2013;50(4):255–63.

    CAS  PubMed  Google Scholar 

  132. Jonker LA, Lebbink CA, Jongmans MCJ, Nievelstein RA, Merks JH, van Dijkum EN, et al. Recommendations on surveillance for differentiated Thyroid carcinoma in children with PTEN hamartoma tumor syndrome. Eur Thyroid J. 2020;9(5):234–42.

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Ngeow J, Eng C. PTEN in hereditary and sporadic Cancer. Cold Spring Harb Perspect Med. 2020;10(4):a036087.

    CAS  PubMed  Google Scholar 

  134. Cameselle-Teijeiro JM, Mete O, Asa SL, LiVolsi V. Inherited follicular epithelial-derived Thyroid carcinomas: from molecular biology to histological correlates. Endocr Pathol. 2021;32(1):77–101.

    PubMed  PubMed Central  Google Scholar 

  135. Griffith CC, Seethala RR. Familial non-medullary thyroid cancer: an update on the genetic and pathologic features. Diagnostic Histopathology. 2016;22(3):101–7.

    Google Scholar 

  136. Ngeow J, Mester J, Rybicki LA, Ni Y, Milas M, Eng C. Incidence and clinical characteristics of thyroid cancer in prospective series of individuals with Cowden and Cowden-like syndrome characterized by germline PTEN, SDH, or KLLN alterations. J Clin Endocrinol Metab. 2011;96(12):E2063–71.

    CAS  PubMed  PubMed Central  Google Scholar 

  137. Ni Y, He X, Chen J, Moline J, Mester J, Orloff MS, et al. Germline SDHx variants modify breast and thyroid cancer risks in Cowden and Cowden-like syndrome via FAD/NAD-dependant destabilization of p53. Hum Mol Genet. 2012;21(2):300–10.

    CAS  PubMed  Google Scholar 

  138. Bennett KL, Mester J, Eng C. Germline epigenetic regulation of KILLIN in Cowden and Cowden-like syndrome. JAMA. 2010;304(24):2724–31.

    CAS  PubMed  PubMed Central  Google Scholar 

  139. Orloff MS, He X, Peterson C, Chen F, Chen J-L, Mester JL, et al. Germline PIK3CA and AKT1 mutations in Cowden and Cowden-like syndromes. Am J Hum Genet. 2013;92(1):76–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  140. Vanhaesebroeck B, Leevers SJ, Ahmadi K, Timms J, Katso R, Driscoll PC, et al. Synthesis and function of 3-phosphorylated inositol lipids. Annu Rev Biochem. 2001;70:535–602.

    CAS  PubMed  Google Scholar 

  141. Yehia L, Niazi F, Ni Y, Ngeow J, Sankunny M, Liu Z, et al. Germline heterozygous variants in SEC23B are associated with Cowden syndrome and enriched in apparently sporadic Thyroid Cancer. Am J Hum Genet. 2015;97(5):661–76.

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Peiretti V, Mussa A, Feyles F, Tuli G, Santanera A, Molinatto C, et al. Thyroid involvement in two patients with Bannayan-Riley-Ruvalcaba syndrome. J Clin Res Pediatr Endocrinol. 2013;5(4):261–5.

    PubMed  PubMed Central  Google Scholar 

  143. Schneider K, Zelley K, Nichols KE, Garber J. Li-Fraumeni Syndrome. In: Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJ, Mirzaa G, et al., editors. GeneReviews® [Internet]. Seattle (WA): University of Washington, Seattle; 1993 [cited 2021 Jun 13]. Available from: http://www.ncbi.nlm.nih.gov/books/NBK1311/

  144. da Cruz Formiga MN, De Andrade KC, Kowalski LP, Achatz MI. Frequency of Thyroid carcinoma in Brazilian TP53 p.R337H carriers with Li Fraumeni syndrome. JAMA Oncol. 2017;3(10):1400–2.

    Google Scholar 

  145. Richards ML. Familial syndromes associated with thyroid cancer in the era of personalized medicine. Thyroid. 2010;20(7):707–13.

    PubMed  Google Scholar 

  146. Yoshida A, Taniguchi S, Hisatome I, Royaux IE, Green ED, Kohn LD, et al. Pendrin is an iodide-specific apical porter responsible for iodide efflux from thyroid cells. J Clin Endocrinol Metab. 2002;87(7):3356–61.

    CAS  PubMed  Google Scholar 

  147. Snabboon T, Plengpanich W, Saengpanich S, Sirisalipoch S, Keelawat S, Sunthornyothin S, et al. Two common and three novel PDS mutations in Thai patients with Pendred syndrome. J Endocrinol Investig. 2007;30(11):907–13.

    CAS  Google Scholar 

  148. Lacka K, Maciejewski A, Stawny B, Lacki JK. Follicular thyroid cancer in a patient with Pendred syndrome. Ann Endocrinol (Paris). 2021;82(6):622.

  149. Tong G-X, Chang Q, Hamele-Bena D, Carew J, Hoffman RS, Nikiforova MN, et al. Targeted next-generation sequencing analysis of a Pendred syndrome-associated Thyroid carcinoma. Endocr Pathol. 2016;27(1):70–5.

    CAS  PubMed  Google Scholar 

  150. Sakurai K, Hata M, Hishinuma A, Ushijima R, Okada A, Taeda Y, et al. Papillary thyroid carcinoma in one of identical twin patients with Pendred syndrome. Endocr J. 2013;60(6):805–11.

    CAS  PubMed  Google Scholar 

  151. Camargo R, Limbert E, Gillam M, Henriques MM, Fernandes C, Catarino AL, et al. Aggressive metastatic follicular thyroid carcinoma with anaplastic transformation arising from a long-standing goiter in a patient with Pendred’s syndrome. Thyroid. 2001;11(10):981–8.

    CAS  PubMed  Google Scholar 

  152. Rothblum-Oviatt C, Wright J, Lefton-Greif MA, McGrath-Morrow SA, Crawford TO, Lederman HM. Ataxia telangiectasia: a review. Orphanet J Rare Dis [Internet]. 2016 Nov 25 [cited 2021 Jun 13];11. Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5123280/

  153. Gu Y, Yu Y, Ai L, Shi J, Liu X, Sun H, et al. Association of the ATM gene polymorphisms with papillary thyroid cancer. Endocrine. 2014;45(3):454–61. https://doi.org/10.1007/s12020-013-0020-1.

    Article  CAS  PubMed  Google Scholar 

  154. Song CM, Kwon T-K, Park BL, Ji YB, Tae K. Single nucleotide polymorphisms of ataxia telangiectasia mutated and the risk of papillary thyroid carcinoma. Environ Mol Mutagen. 2015;56(1):70–6.

    CAS  PubMed  Google Scholar 

  155. Ulusoy E, Edeer-Karaca N, Özen S, Ertan Y, Gökşen D, Aksu G, et al. An unusual manifestation: papillary thyroid carcinoma in a patient with ataxia-telengiectasia. Turk J Pediatr. 2016;58(4):442–5.

    PubMed  Google Scholar 

  156. Maillard S, Damiola F, Clero E, Pertesi M, Robinot N, Rachédi F, et al. Common variants at 9q22.33, 14q13.3, and ATM loci, and risk of differentiated thyroid cancer in the French Polynesian population. PLoS One. 2015;10(4):e0123700.

    PubMed  PubMed Central  Google Scholar 

  157. Geoffroy-Perez B, Janin N, Ossian K, Laugé A, Croquette MF, Griscelli C, et al. Cancer risk in heterozygotes for ataxia-telangiectasia. Int J Cancer. 2001;93(2):288–93.

    CAS  PubMed  Google Scholar 

  158. Dombernowsky SL, Weischer M, Allin KH, Bojesen SE, Tybjaerg-Hansen A, Nordestgaard BG. Risk of cancer by ATM missense mutations in the general population. J Clin Oncol. 2008;26(18):3057–62.

    CAS  PubMed  Google Scholar 

  159. Yalçin S, Kirli E, Ciftci AO, Karnak I, Resta N, Bagnulo R, et al. The association of adrenocortical carcinoma and thyroid cancer in a child with Peutz-Jeghers syndrome. J Pediatr Surg. 2011;46(3):570–3.

    PubMed  Google Scholar 

  160. Kopacova M, Tacheci I, Rejchrt S, Bures J. Peutz-Jeghers syndrome: diagnostic and therapeutic approach. World J Gastroenterol. 2009;15(43):5397–408.

    PubMed  PubMed Central  Google Scholar 

  161. Wei S, LiVolsi VA, Brose MS, Montone KT, Morrissette JJD, Baloch ZW. STK11 mutation identified in Thyroid carcinoma. Endocr Pathol. 2016;27(1):65–9.

    CAS  PubMed  Google Scholar 

  162. Buryk MA, Picarsic JL, Creary SE, Shaw PH, Simons JP, Deutsch M, et al. Identification of unique, heterozygous germline mutation, STK11 (p.F354L), in a child with an encapsulated follicular variant of papillary Thyroid carcinoma within six months of completing treatment for neuroblastoma. Pediatr Dev Pathol. 2015;18(4):318–23.

    PubMed  Google Scholar 

  163. Hagelstein-Rotman M, Meier ME, Majoor BCJ, Cleven AHG, Dijkstra PDS, Hamdy NA, et al. Increased prevalence of malignancies in fibrous dysplasia/McCune-Albright syndrome (FD/MAS): data from a National Referral Center and the Dutch National Pathology Registry (PALGA). Calcif Tissue Int. 2021;108(3):346–53.

    CAS  PubMed  Google Scholar 

  164. Haddad RI, Nasr C, Bischoff L, Busaidy NL, Byrd D, Callender G, et al. NCCN guidelines insights: Thyroid carcinoma, version 2.2018. J Natl Compr Cancer Netw. 2018;16(12):1429–40. https://doi.org/10.6004/jnccn.2018.0089.

    Article  Google Scholar 

  165. Haugen BR, Alexander EK, Bible KC, Doherty GM, Mandel SJ, Nikiforov YE, et al. 2015 American Thyroid Association management guidelines for adult patients with Thyroid nodules and differentiated Thyroid Cancer: the American Thyroid Association guidelines task force on Thyroid nodules and differentiated Thyroid Cancer. Thyroid. 2016;26(1):1–133.

    PubMed  PubMed Central  Google Scholar 

  166. Filetti S, Durante C, Hartl D, Leboulleux S, Locati LD, Newbold K, et al. Thyroid cancer: ESMO clinical practice guidelines for diagnosis, treatment and follow-up†. Ann Oncol. 2019;30(12):1856–83.

    CAS  PubMed  Google Scholar 

  167. Takahashi M, Saenko VA, Rogounovitch TI, Kawaguchi T, Drozd VM, Takigawa-Imamura H, et al. The FOXE1 locus is a major genetic determinant for radiation-related thyroid carcinoma in Chernobyl. Hum Mol Genet. 2010;19(12):2516–23.

    CAS  PubMed  Google Scholar 

  168. Mussazhanova Z, Rogounovitch TI, Saenko VA, Krykpayeva A, Espenbetova M, Azizov B, et al. The Contribution of Genetic Variants to the Risk of Papillary Thyroid Carcinoma in the Kazakh Population: Study of Common Single Nucleotide Polymorphisms and Their Clinicopathological Correlations. Front Endocrinol (Lausanne) [Internet]. 2021 [cited 2021 Jun 8];11. Available from: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7862756/

Download references

Acknowledgements

Not applicable.

Funding

No specific funding used for this manuscript.

Author information

Authors and Affiliations

Authors

Contributions

All persons who meet authorship criteria are listed as authors, and all authors certify that they have participated sufficiently in the work to take public responsibility for the content, including participation in the concept, design, analysis, writing, or revision of the manuscript.

Corresponding author

Correspondence to Mohammad R. Akbari.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The author declares that they have no relevant or material financial interests that relate to the research described in this paper.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kamani, T., Charkhchi, P., Zahedi, A. et al. Genetic susceptibility to hereditary non-medullary thyroid cancer. Hered Cancer Clin Pract 20, 9 (2022). https://doi.org/10.1186/s13053-022-00215-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13053-022-00215-3

Keywords