Sie sind auf Seite 1von 21

Molecular and Cellular Endocrinology 322 (2010) 8–28

Contents lists available at ScienceDirect

Molecular and Cellular Endocrinology


journal homepage: www.elsevier.com/locate/mce

Review

Molecular prognostic markers in papillary and follicular thyroid cancer: Current


status and future directions
Daria Handkiewicz-Junak a , Agnieszka Czarniecka b , Barbara Jarzab
˛ a,∗
a
Department of Nuclear Medicine and Endocrine Oncology, Maria Sklodowska-Curie Memorial Cancer Center and Institute of Oncology, Gliwice Branch, Wybrzeze Armii Krajowej
14, 44-100 Gliwice, Poland
b
Department of Oncologic Surgery, Maria Sklodowska-Curie Memorial Cancer Center and Institute of Oncology, Gliwice Branch, Wybrzeze Armii Krajowej 14, 44-100 Gliwice, Poland

a r t i c l e i n f o a b s t r a c t

Article history: Gene expression profiling shows that, by gene signature, the difference between BRAF-positive and BRAF-
Received 7 July 2009 negative PTC is so distinct that BRAF-positive cancer may be regarded as a molecular subtype of papillary
Received in revised form thyroid cancer (PTC). Since much enthusiasm surrounds the BRAF-oncogene as a molecular prognostic
20 December 2009
factor, a central focus of our consideration is to weigh the current arguments for and against applying BRAF
Accepted 9 January 2010
mutation status of the tumor in clinical practice. The frequency of BRAF mutation in PTC is high—45% on
average, with values over 70–80% in some populations. This will mean that implementing BRAF mutation
Keywords:
as a factor of poor prognosis will shift many PTC patients, considered up to now as low risk ones, to the
Papillary thyroid cancer
Follicular thyroid cancer
more extensive treatment. We estimate that 31% of all PTC patients and 39% of those diagnosed with
Prognosis stage I–II disease will face the risk of overtreatment if the decision will be based on the BRAF-positivity
BRAF of their tumors. Also, the risk of undertreatment in the young patients with BRAF-negative tumors is
Molecular markers evaluated with 26%. We think that, as of now, the evidence-based support for such consequences is
still weak. Thus, there is urgent need to look for genes or gene signatures which will be helpful in the
stratification of BRAF-positive tumors to specify these with poor prognosis with higher accuracy, needed
for clinical decisions. Considering this, in the review we summarize the present status of knowledge on
other prognosis-related gene expression changes in papillary and follicular cancer and relate them to he
tumor’s biology.
© 2010 Elsevier Ireland Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2. Pathological and clinical basis for evaluating prognosis in papillary and follicular thyroid cancers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1. Histopathological and clinical characteristics of DTCs and their influence on disease prognosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2. Clinicopathological staging of DTC and its limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3. Molecular mechanisms of neoplastic transformation in papillary and follicular thyroid cancer: prognosis-related aspects . . . . . . . . . . . . . . 11
2.3.1. Activation of the RAS–RAF–MEK–ERK1/2 pathway and the PI3K/AKT pathway in thyroid cancers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3. Prognostic impact of the inititiating molecular events in papillary and follicular thyroid cancers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.1. The role of BRAF proto-oncogene in PTC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.1.1. BRAF mutation is the most frequent initiating molecular event in PTC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.1.2. BRAF mutations are associated with poorer prognosis of PTC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2. Is it the time to apply BRAF as clinically relevant prognostic/predictive factor in PTC? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.3. Other molecular initiating events specific for PTC—do they have prognostic potential? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3.1. RET/PTC rearrangements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3.2. NTRK rearrangements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3.3. Gene rearrangements are frequent in thyroid cancers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.4. Is there some multiclonality in papillary thyroid cancer? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.5. Gene mutations important for development of both papillary and follicular thyroid cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.5.1. RAS mutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

∗ Corresponding author. Tel.: +48 322789339; fax: +48 322789325.


˛
E-mail addresses: dhandkiewicz@io.gliwice.pl (D. Handkiewicz-Junak), aczarniecka@io.gliwice.pl (A. Czarniecka), bjarzab@io.gliwice, bjarzab@io.gliwice.pl (B. Jarzab).

0303-7207/$ – see front matter © 2010 Elsevier Ireland Ltd. All rights reserved.
doi:10.1016/j.mce.2010.01.007
D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 9

3.6. Follicular thyroid cancer-specific gene mutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17


3.6.1. PAX8/PPAR rearrangements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.6.2. RAS homolog 1 (ARH1) silencing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.7. Mutations of other genes and their importance for thyroid cancer and its outcome . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.7.1. PIK3CA mutations and amplification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.7.2. PTEN mutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.7.3. ␤-Catenin mutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.7.4. TP53 mutations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4. Prognostic aspects of other gene expression changes in papillary and follicular thyroid cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.1. Single genes expression changes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.1.1. MET proto-oncogene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.1.2. Epithelial growth factor receptor (EGFR). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.1.3. Mitogen-inducible gene-6 (MIG-6) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1.4. Vascular epithelial growth factor (VEGF) and VEGF receptor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1.5. E-cadherin and catenins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1.6. Cyclin D1 and other cell cycle regulators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1.7. Mucin 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1.8. S100A4 calcium-binding protein . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1.9. Osteopontin and CD44v6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1.10. Immunity-related genes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.2. Prognosis-related gene signatures in thyroid cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2.1. Gene expression profiling and prognosis of thyroid cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2.2. Metastasis-associated genes in thyroid cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2.3. Epithelial-to-mesenchymal transition in PTC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.3. Genetic germline background and the prognosis of papillary and follicular thyroid cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
5. Final considerations on prognostic impact of genetic factors in papillary and follicular thyroid cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
5.1. Relation between BRAF status and gene expression profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
5.1.1. Is BRAF-positive DTC a distinct molecular subtype? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
5.1.2. BRAF-mutated PTCs exhibit fewer thyroid-specific gene expression features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5.1.3. BRAF mutation is associated with the more aggressive tumor phenotype in PTC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5.2. Major questions in the therapeutic strategy in DTC which can be solved by molecular diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5.3. Conclusions from the transcriptome-wide analyses of DTCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

1. Introduction sion into two molecular subtypes – luminal and basal – is slowly
gaining influence in formulating disease management strategies
Papillary thyroid cancer (PTC) and follicular thyroid cancer (Sorlie et al., 2001; van’t Veer et al., 2005).
(FTC) are both described clinically as differentiated thyroid cancers In the present review, we highlight that the molecular diag-
(DTCs) due to their tumor cells structurally and functionally resem- nosis of thyroid cancer is already possible (Eszlinger et al., 2008;
bling normal follicular cells and due to their sharing a relatively Giordano, 2008). We therefore examine whether, to improve prog-
indolent natural history and good responsiveness to surgery and nostic/predictive ability, molecular diagnosis should now supplant
radioiodine in most – but far from all – cases. DTC constitutes the or at least supplement the clinicohistopathological evaluation of
commonest endocrine malignancy, and among neoplastic diseases, DTC. This question is increasingly pertinent, as in recent years,
possesses the most favorable overall prognoses. However, over 10% both research on single, cancer-specific mutations and genome-
of patients eventually die of DTC and an even greater proportion wide DNA microarray-based approaches have added greatly to our
faces the morbidity of recurrences (Eustatia-Rutten et al., 2006; understanding of DTC biology and clinical behavior. In our discus-
Mazzaferri and Kloos, 2001; Sherman, 2003). In these respects, lit- sion of the prognostic relevance of molecular studies, we focus on
tle to nothing has improved in DTC management in the last twenty data regarding gene mutations, essential for neoplastic transforma-
years (Clark and Duh, 1991). tion of follicular cells and, then, for gene expression changes, judged
As in other cancers, numerous studies in DTC have concentrated both by RNA and protein levels. We present details of microarray-
on relatively simple clinicopathological variables, e.g., primary based comparative genome hybridisation (CGH) studies on gene
tumor size, extent of disease, pathological histotype, patient’s age amplifications/deletions and of studies on epigenomic changes
at diagnosis or sex, to formulate risk group stratification or stag- insofar as this information is germane to the current role and
ing systems. The goal has been to select patients who will do well future investigation of molecular prognostic factors in DTC. Since
without further therapy (i.e., to identify prognostic factors) or who the BRAF-oncogene (v-raf murine sarcoma viral oncogene homolog
will do well with some treatments but not others (i.e., to identify B1) is the major genetic factor in PTC (Puxeddu and Moretti, 2007;
predictive factors). Xing, 2007), an important focus of the present review is to weigh
In recent years, the impact of molecular diagnosis is becoming the current arguments for and against applying BRAF mutation sta-
even more substantial in cancer care. This trend is already evident in tus in clinical practice. As we discuss in more detail below, the data
many hematological malignancies, where the presence of particu- on the correlation of BRAF status with DTC clinicopathological fac-
lar chromosomal rearrangement influences the natural history and tors are substantial and even show an impact on survival. However,
the choice of treatment, i.e., is not only of prognostic significance from the clinical point of view, a factor present in 50% of cases is of
but also of predictive significance (Haferlach et al., 2003; Jabbour limited use in managing a disease which has a poor outcome in no
et al., 2008; Meijerink et al., 2009). A similar phenomenon is also more than 10–15% patients, thus, a more detailed stratification is
occurring with solid tumors such as breast cancer, where the divi- necessary and we look for this possibility.
10 D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28

Our review begins by summarizing the current understanding et al., 2005; Tubiana et al., 1985; Zaydfudim et al., 2008). In fact,
of histopathological and clinical prognostic factors in DTC. Next, in the past, lymph-node involvement was even considered to be
we describe BRAF mutations and other molecular events occurring a favorable prognostic sign, until the effect of age was recog-
early in the thyroid neoplastic transformation in the context of their nized: children and young adults with PTC have both far more
relationship to thyroid cancer prognosis. Subsequently, we discuss frequent lymph-node involvement and a far better outcome than
the expression changes in other single genes other than BRAF that do older patients. However, almost all studies unequivocally agree
are known to have prognostic utility in DTC. Lastly, we present the that lymph-node involvement is associated with a higher risk of
status of relevant genome-wide analyses and provide some closing locoregional recurrences (Beasley et al., 2002; Handkiewicz-Junak
remarks where we consider not only the prognostic value of genetic et al., 2007; Harwood et al., 1978; Sakorafas et al., 2009).
markers but also comment briefly on their predictive contribution.
2.2. Clinicopathological staging of DTC and its limitations

2. Pathological and clinical basis for evaluating prognosis At least 18 staging systems for PTC, FTC (Akslen, 1993; Cady
in papillary and follicular thyroid cancers and Rossi, 1988; DeGroot et al., 1990; Greene et al., 2002; Hay et
al., 1987, 1993; Mazzaferri and Jhiang, 1994; Pasieka et al., 1992;
2.1. Histopathological and clinical characteristics of DTCs and Shaha et al., 1994; Sherman et al., 1998) or both, have been formu-
their influence on disease prognosis lated based on retrospective patient outcome evaluations (Lang et
al., 2007a,b). In Europe, the UICC/TNM system is the most widely
The follicular cells of the thyroid gland give rise to a variety adopted. It considers three main factors:
of tumors that differ markedly in their morphology and biologi-
cal and clinical behavior. Based on a number of diagnostic criteria • T status—primary tumor size and extrathyroidal extension,
and morphological hallmarks, these tumors may be divided into • N status—presence or absence of lymph node metastases,
five major types: papillary, follicular, oncocytic, poorly differenti- • M status—presence or absence of distant metastases.
ated, and undifferentiated (anaplastic), the first two of which in
turn may be subdivided into a number of variants (LiVolsi and Asa, The UICC/TNM system also considers patient age at diagnosis,
1994; LiVolsi and Baloch, 2004). By far the most common is PTC, a factor not commonly used in other non-TNM-based DTC staging
representing up to 80% of all thyroid malignancies. This prevalence systems. All patients below 45 years of age who are diagnosed with-
is distantly followed by that of FTC, ranging from less than 10% or out distant metastases, are considered to have stage I disease, i.e.,
15% to not more than 25%, depending in the series, while oncocytic belong to the “low-risk” group.
histotype (either variant of PTC or FTC) constitutes not more than 2% The clinicopathological factors upon which the majority of
of thyroid cancers (Kushchayeva et al., 2008). Several histopatho- staging systems are based are summarized in Table 1. These vari-
logical features of DTC have been claimed to influence its prognosis. ables frequently overlap, e.g., older or very young patients tend
In PTC, the tall cell variant has been shown to have a worse prog- to more often have lymph node or distant metastases or both;
nosis than does the classical type (Ito et al., 2008). Concomitant young patients with lymph node metastases more often suffer
thyroiditis, diffuse lymphocytic infiltration, or both may favorably from distant metastases (Jarzab et al., 2005a). Such overlap is usu-
influence the course of PTC as well as that of FTC (Kashima et al., ally addressed with multivariate analysis, which can help select
1998; Loh et al., 1999; Modi et al., 2003). independent variables while rejecting factors that are only surro-
FTC, especially if widely invasive or oncocytic, was claimed to gates for those variables. Nonetheless, even with that statistical
have a worse prognosis than does PTC (Gulcelik et al., 2007; Passler approach, results are sometimes contradictory, as in the evaluation
et al., 2004). However, a recent study of more than 1000 patients of tumor histopathology (Gulcelik et al., 2007; Verburg et al., 2009)
found no difference in cancer-specific survival between PTC and or of lymph node metastases (Bardet et al., 2008; McConahey et al.,
FTC when one accounted for clinical prognostic factors such as age 1986). In the case of histopathology, the contradictory results may
at diagnosis, primary tumor size and the presence of extrathyroidal be attributable to the somewhat subjective, not fully standardized
invasion or of metastases (Verburg et al., 2009). In FTC, the prog- method of specimen evaluation. The rate of disagreement during
nosis for the minimally invasive subtype is excellent (Thompson DTC evaluation can be high, especially in FTC where it can reach up
et al., 2001). Fig. 1 compares a variety of commonly used clin- to 50% of cases (Franc et al., 2003; Lange et al., 2006).
ical prognostic factors with respect to the associated significant Table 2 provides the survival by disease stage or risk group
increases in the risks of death or DTC recurrence that were found according to the most popular staging systems. As seen in Table 2,
in a recent large, long-term study of DTC patients conducted at our the systems show marked survival differences, especially for “high-
center (Czarniecka, 2004). As seen in that figure, the presence of risk” groups/advanced stages, in which, according to most systems,
distant metastases is the strongest poor prognostic factor for over- more than 50% of patients eventually succumb to disease (Durante
all survival in DTC (Elisei et al., 2008; Mazzaferri and Jhiang, 1994; et al., 2006; Sampson et al., 2007; Schlumberger et al., 1996). It
Mazzaferri and Kloos, 2001). However, there are exceptions to this is however, not established which patients in less advanced sub-
rule, as distant metastases in young patients are not regarded as a groups do or do not require more intensive treatment to cure their
poor prognostic factor. In fact, younger patients with distant metas- cancer. This results in a very wide “gray zone” of DTC prognosis,
tases have a better prognosis than do older patients diagnosed with which leads to a significant risk of over- or undertreatment in a sub-
locally invasive tumors, lymph node metastases, or both (Jarzab et stantial number of patients (Gulcelik et al., 2007; Tuttle et al., 2008).
al., 2005a). The MACIS and TNM staging systems have been demonstrated to be
The prognostic impact of lymph-node involvement in DTC has the most accurate prognostic systems in some (Lang et al., 2007a,b;
long been a matter of controversy, especially regarding survival. Passler et al., 2003) but not all studies (Brierley et al., 1997). Never-
Several studies failed to demonstrate an influence on mortality theless, even if evaluated with the best score, they explained barely
rates (DeGroot et al., 1990; Mazzaferri and Young, 1981) while oth- about 20% of the observed variation in survival time (Lang et al.,
ers have identified an association between lymph node metastases 2007a,b) and only a few studies yielded better results (D’Avanzo et
and an increased risk of cancer-related death, especially when the al., 2004).
involved lymph nodes are large, bilateral or located in the medi- Exacerbating this problem, the extent of treatment (e.g., rad-
astinum (Czarniecka, 2004; Mazzaferri and Kloos, 2001; Podnos icalness of surgery or use or activity of radioiodine) that is
D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 11

Fig. 1. Clinical and pathological risk factors of overall and recurrence-free survival in thyroid cancer (multivariate regression analysis of 1141 DTC patients, followed-up by
5 years) (Czarniecka, 2004).

recommended in guidelines is not consistent between “high-risk” commonalities in the clinical course of both cancers and cre-
or “low-risk” groups according to different staging systems. It is, of ates a rationale for similar therapeutic approaches to both of
course, well known that the extent of treatment can influence dis- them.
ease outcome (Czarniecka, 2004; Gulcelik et al., 2007; Lang et al.,
2007a,b; Passler et al., 2003). In addition, since almost all current 2.3.1. Activation of the RAS–RAF–MEK–ERK1/2 pathway and the
clinicohistopathological prognostic factors are based on postsur- PI3K/AKT pathway in thyroid cancers
gical tumor and patient evaluation, none of these staging systems DTC harbors several highly prevalent genetic alterations, some
can be applied pre-operatively, and hence help tailor the extent of of which are seen only in this cancer or are characteristic of one of its
surgery. histotypes. In PTC, the most important oncogenic mechanism, seen
In summary, then, the utility of the current clinicohistopatho- in about 80% of tumors, is the activation of the mitogen-activated
logical staging systems for DTC is limited by the abundance protein kinase (MAPK) pathway (Fagin, 2004; Fagin and Mitsiades,
of systems, their frequent overlap and sometimes contradictory 2008), which once constitutively activated, leads to tumorigene-
results, their far from perfect match with short- or long-term prog- sis (Hilger et al., 2002; Peyssonnaux and Eychene, 2001). Three
nosis, and their inapplicability pre-operatively. most important initiating events, RET/PTC (rearranged during trans-
fection/papillary thyroid cancer), RAS (resistance to audiogenic
2.3. Molecular mechanisms of neoplastic transformation in seizures) and BRAF mutation, are regarded as mutually exclusive,
papillary and follicular thyroid cancer: prognosis-related aspects alternative triggers for the activation of the pathway (Fagin, 2004).
The mutual exclusivity between BRAF and the RAS mutation is char-
From the molecular point of view, papillary and follicular thy- acteristic not only of thyroid cancer but also of several other human
roid cancers are regarded as different diseases. This opinion is cancers (Xing, 2005). The activating BRAF mutation has only been
supported by the disparate molecular initiating events leading to recognized relatively recently (Davies et al., 2002), but in fact is the
neoplastic transformation. Also supporting this perspective are the most frequent trigger in PTC.
differences in DNA ploidy level (PTCs are generally diploid, FTCs A single oncogenic alteration along the receptor tyrosine
aneuploid) and the differences in localization and in the number kinase–RAS–RAF–MEK–ERK1/2 pathway is likely sufficient to drive
of gains and losses seen in comparative genomic hybridisation thyroid cell neoplastic transformation, although further supportive
(CGH)-based analysis. In PTC, aneuploidy, although observed in molecular events are necessary. BRAF mutation and RET/PTC (and
minority of tumors (Rodrigues et al., 2007), is significantly associ- NTRK1—neurotrophic tyrosine kinase, receptor, type 1) rearrange-
ated with cancer-specific death (Sturgis et al., 1999). The profound ments differ to some extent in their effects on the shared oncogenic
differences in gene expression will be described below. On the pathway, respectively, resulting more frequently in the classic vari-
other hand, many subsequent down-stream molecular mecha- ant or the solid variant of PTC, while RAS mutations and RASSF1A
nisms are very similar between PTC and FTC, which explains the (Ras association (RalGDS/AF-6) domain family member 1) methy-
12 D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28

Table 1
Clinical prognostic factors used in different staging systems for DTC.

Factor Positive prognostic indicator Negative prognostic indicator

Age Younger (usually below 40–45 years of age) Older


Primary tumor—diameter Small (the best prognosis for tumors ≤1 cm) Large
Primary tumor—extension beyond thyroid capsule No Yes
Lymph node metastasis Absent Present
Distant metastases Absent Present
Surgical resection Total thyroidectomy and appropriate extent of Incomplete removal of cancer tissues and/or
lymphadenectomy ensuring complete removal lack of total thyroidectomy and of appropriate
of cancer tissues extent of lymphadenectomy

Table 2
Survival by disease stage or risk group in selected DTC staging systems.

Staging system DTC type (number of patients) Primary outcome end-point Disease stage or risk group

TNM stage (Greene et al., 2002) PTC (6590) 5-Year survival I II 93% III IV
97% 84% 39%
FTC (1129) 5-Year survival 95% 90% 69% 41%

MACIS (Hay et al., 1993) PTC (1779) 20-Cause-specific survival rates I II 89% III IV
99% 56% 24%
Ohio State University Staging PTC and FTC (1355) Cancer-related deaths I II 6% III IV
system (Mazzaferri and 0% 15% 65%
Jhiang, 1994)
AMES (Cady and Rossi, 1988) PTC and FTC (821) Cancer-related deaths Low High
1.8% 46%
NTCTCS system (Sherman et PTC and FTC (1607) 5-Year CSS I II 100% III IV
al., 1998) 99.8% 91.9% 48.9%
University of Chicago (DeGroot PTC (269) 10-Year CSS I II 100% III IV
et al., 1990) 100% 87% 35%
GAMES Risk Groups (Shaha et PTC and FTC (1038) 10-Year survival Low Intermediate85% High
al., 1994) 99% 57%
AGES (Hay et al., 1987) PTC (860) 25-Year CSS Low High
98% 65%
DAMES (Pasieka et al., 1992) PTC (74) Cancer-related death or recurrence Low Intermediate45% High
8% 100%
SAG (Akslen, 1993) PTC (173) 10-Year CSS I II 88% III
98.3% 39%

CSS, cancer-specific survival; DTC, differentiated thyroid cancer; FTC, follicular thyroid cancer; PTC, papillary thyroid cancer.

lation are more prone to induce the follicular variant of PTC (Xing, thyroid cancer-specific. To the contrary, BRAF belongs to the com-
2005). monest oncogenes, mutated in 7–9% of all malignant solid tumors,
MEK–ERK1/2 pathway activation also plays a significant role with the highest prevalence (about 60%) in melanomas (DeLuca
in FTC tumorigenesis (Liu et al., 2008). However, in FTC, neither et al., 2008), and a frequent presence in ovarian, colorectal, and
RET/PTC nor the BRAF mutation can be found, while RAS mutations lung cancers and even some low-grade astrocytomas (Pfister et al.,
are quite frequent, seen not only in follicular cancers, but also in 2008). Interestingly, although the activation of the MAPK signaling
follicular adenomas (FA). This pattern is repeated in the case of pathway is regarded as frequent in human cancer, the evidence for
PAX8/PPAR rearrangements, which however do not directly lead this is most compelling in thyroid cancer and melanoma (Johansson
to MEK–ERK activation and cause different down-stream effects et al., 2007; Knauf and Fagin, 2009).
(Giordano et al., 2006).
The phosphatidylinositol 3-kinase (PI3K)/Akt pathway also
plays a fundamental role in the regulation both of normal cell 3.1.1. BRAF mutation is the most frequent initiating molecular
growth, mitosis and survival and, when disarranged, of tumorige- event in PTC
nesis. PI3K/AKT alterations are found in every histotype of thyroid After melanoma, PTC is the second human malignancy where
cancer, frequently in FTC and, even more distinctly, in ATC. In FTC, BRAF mutations are especially frequent, found in 29–87% (with a
phosphorylation of AKT, the key player in this pathway, was far mean frequency of about 45%) of cases (Kim et al., 2009; Kimura
more frequent than that of ERK (MAPK kinase effector) (Liu et al., et al., 2003; Puxeddu et al., 2004; Xing, 2007). BRAF mutations are
2008). not found exclusively in PTC but also, albeit less often (not more
than 20–26% of cases, mainly those originating from PTC) in ATC
(Nikiforova et al., 2003a; Smallridge et al., 2009; Takano et al.,
3. Prognostic impact of the inititiating molecular events in 2007).
papillary and follicular thyroid cancers Although more than 40 mutations have been identified in the
BRAF gene, the most significant hot spot mutation, accounting for
3.1. The role of BRAF proto-oncogene in PTC over 90% of all BRAF mutations, is a thymidine to adenine transver-
sion at nucleotide 1799 (T1799A) in exon 15. This substitution
BRAF is a serine–threonine kinase, expressed in a tissue-specific leads to a valine-to-glutamate transversion at residue 600 near
manner and abundant in thyroid follicular cells, which among the the catalytic center of the protein (BRAFV600E) and is believed to
three known RAF kinases, most potently activates the MAPK path- produce a constitutively active kinase by disrupting hydrophobic
way (Peyssonnaux and Eychene, 2001) (for reviews see Puxeddu interactions between residues in the activation loop and residues
and Moretti, 2007; Xing, 2005). Activating BRAF mutations are not in the adenosine triphosphate (ATP) binding site (Wan et al., 2004).
D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 13

BRAFV600E potentiates the catalytic activity of the enzyme by more Elisei et al., 2008; Guan et al., 2009; Kebebew et al., 2007; Lee et al.,
than 500 times (Kim et al., 2005). 2009; Rosenbaum et al., 2005) were based on univariate analyses
BRAF mutation is an early event in PTC tumorigenesis, which that could not distinguish between overlapping significant factors.
can be found even in minute PTCs, less than 4 mm in diameter In that case what seems to be BRAF related in univariate analysis can
(Ugolini et al., 2007). The constitutively active kinase subsequently result from other unknown factors affecting studied outcome (e.g.,
leads to tumorigenesis through aberrant activation of the MAPK disease stage or recurrences) and be correlated with BRAF status.
pathway (Davies et al., 2002; Garnett and Marais, 2004). Such acti- For example, BRAF mutations are relatively rare in PTC cases with
vation includes hyperphosphorylation of retinoblastoma protein concommitant autoimmune thyroiditis (Nikiforova et al., 2002),
(RB), which releases inhibition of E2F-dependent transcription fac- and here the immune response may influence the clinical course
tors, allowing the cell to pass from G1 into S phase, increasing more than do the molecular consequences of initiating mutation
growth and promoting survival (for a review of this process see itself. Although the multivariate analysis may be somewhat help-
Knauf and Fagin, 2009). At the same time, apoptosis is inhibited, ful in making the distinction between diverse confounding factors,
probably due to the action of NF␬B transcription factor (Palona it may only solve some but not all the problems, as only a limited
et al., 2006). Transgenic mouse studies clearly demonstrated the number of factors can be included and other, theoretically deci-
driving force of the BRAF mutation in promoting malignant trans- sive factors in the outcome of the disease can be unintentionally
formation of thyrocytes and extrathyroidal invasion of PTC (Knauf omitted. Nevertheless at present, multivariate analysis seems to be
et al., 2005). One important molecular mechanism involved in this the best way to obtain the least biased results. A good example is
process is BRAF mutation-promoted methylation and hence silenc- the already discussed study by Lupi et al., where BRAF mutation
ing of the tissue inhibitor of metalloproteinase-3 (TIMP-3; Hu et al., positively correlated with extrathyroidal extension in univariate
2004) and resultant overexpression of metalloproteinases (Melillo analysis, but was abolished by the status of thyroid capsule in mul-
et al., 2005; Mesa et al., 2006; Palona et al., 2006). tivariate approach.
The BRAF-oncogene induces chromosomal instability A second issue in interpreting these studies is how precisely
(Mitsutake et al., 2005), and is it not only an initiator of PTC clinicopathological factors are defined. For example, both mini-
but is necessary for the maintenance of PTC proliferation and mal invasion beyond the thyroid capsule and gross invasion of
tumorigenicity (Liu et al., 2007b). muscles and other neck tissue can be regarded as extrathyroidal
disease extension, but their clinical significance differs substan-
3.1.2. BRAF mutations are associated with poorer prognosis of tially: tumors with gross invasion into surrounding tissues have a
PTC worse prognosis and a higher risk for recurrence. Differences in the
Based on molecular results, numerous studies have investigated methodology of BRAF detection can also cause conflicting results
the clinical significance of BRAF mutation in PTC. The general ten- among studies.
dency of the studies to show the association of this mutation with Surprisingly, distant metastases, the most significant factor of
factors related to poor prognosis is rather convincing (Table 3 ). poor PTC outcome, were not associated with BRAFV600E mutation
However, a careful overview of the literature indicates that the clin- as a single factor. However, when that mutation was associated
ical data on the link between BRAF mutation and DTC outcome are with other genetic events, the conjunction of events appeared as
not very consistent. The first data were summed up in a large meta- an independent predictor of distant metastasis (Costa et al., 2008).
analysis of 1168 patients and showed that BRAF mutation correlated A more definitive assessment of the impact of BRAF status on
with histologic subtype, presence of extrathyroidal extension, and PTC prognosis can be derived solely from studies evaluating the
advanced clinical stages, but not with age, sex, race, or tumor size of long-term outcome of the disease. Only recently have reports on
PTC patients (Lee et al., 2007). More recent studies addressed even disease-free survival or overall survival in BRAF-positive versus
more carefully the correlation between BRAF status and PTC clini- BRAF-negative PTC patients been published (Table 4). Ito et al.
copathological factors: two such studies (Frasca et al., 2008; Wang (2009) evaluated disease-free survival and distant metastases-free
et al., 2008a), which altogether included more than 400 patients, survival with respect to BRAF mutation status in more than 600
found in multivariate analysis a positive correlation between BRAF patients followed for a mean 83 ± 35 months. The BRAFV600E muta-
mutation and extrathyroidal extension, while two other such stud- tion was detected in 38% of their series, but these investigators
ies totaling about twice as many patients did not (Kebebew et al., did not search for other far less frequent BRAF point mutations or
2007; Lupi et al., 2007). In the analysis of Lupi et al., encompassing rearrangements. The only significant difference among patient sub-
500 consecutive PTC patients at the University of Pisa, the one-way groups was the frequency of BRAF mutations in micro-PTC (28%)
univariate correlations of BRAF mutation positivity with extrathy- versus tumors larger than 1 cm (41%) (P = 0.0175), however, BRAF
roidal extension and tumor stage-related data were very distinct. mutation prevalence did not successively increase with size in
However, the multivariate analysis discarded all the factors except patients having tumors larger than 1 cm. In multivariate analysis,
lack of a tumor capsule (whether the tumor was single or multifo- BRAF mutation was not linked to patient gender, disease stage, mas-
cal) (P = 0.0005 in multivariate analysis). In the Lupi study, lack of sive extrathyroidal extension, or lymph node/distant metastases at
encapsulation remained statistically significant in separate multi- diagnosis. Importantly, the authors also did not find any difference
variate analyses for micro-PTC, larger tumors and follicular variant in 5- or 10-year disease-free survival or metastases-free survival,
PTC but not for classical variant PTC. Presence of tumor capsule, which were about 90% in both BRAF-positive or BRAF-negative
although relatively rare in PTC (present in 21.7% of the Lupi series), cases. This observation confirmed the findings of a previous mul-
is a strong predictor of an especially good disease-free prognosis, ticenter Italian report involving PTC of all types (Fugazzola et al.,
seen very distinctly in follicular variant PTC (Kakudo et al., 2004). 2004) and a Korean study (Kim et al., 2006b) involving the clas-
Summing up the analysis of the relationship of BRAF muta- sic PTC variant. Contrary conclusions, supporting the correlation of
tions with clinicopathological factors, there is only one factor BRAF status with disease-free survival, were reported by Kebebew
shown by the majority of studies to correlate with BRAF et al. (2007) (mean follow-up of 6 years) who in multivariate
mutations—extrathyroidal tumor extension, which is however analysis, noted a significant association of BRAF mutation with
questioned if tumor encapsulation is also included in the analy- recurrence risk.
sis. For one perusing these studies, questions arise on the origins of The recent study by Elisei et al. (2008), who analyzed 102
the discrepancies between the study results. Many studies on the patients with a median follow-up of >10 years, is the only published
correlation of BRAF with clinicopathological data (Costa et al., 2008; study until now to address the relationship of BRAF mutation status
14
Table 3
Correlation of BRAF mutation and clinicopathological prognostic factors in papillary thyroid carcinoma: data from recent studies.

D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28


Author Number of Type of statistical Number of Tumor size Gender Age Extrathyroidal Tumor Lymph node Distant Cancer stage
patients analysis BRAF-positive cases extension capsule metastases metastases
(%)
Rosenbaum et al. (2005) 85 Univariate 54 (64%) – – 0.0001 – – – – –

Kim et al. (2006a) 103 Univariate 34 (33%) 0.0001 ns 0.01 – – 0.0001 – 0.0001
Multivariate ns – ns – – 0.0001 – ns

Kim et al. (2006b) 203 Univariate 149 (73%) 0.006 0.006 ns 0.06 – ns – ns
Multivariate 0.03 0.04 0.005 (gross invasion) – 0.02 – –

Kebebew et al. (2007) 274 Univariate 133 (49%) ns ns 0.03 ns ns ns ns 0.04

Lupi et al. (2007) 500 Univariate 217 (44%) – ns ns 0.0001 <0.0001 0.0009 – 0.00001
Multivariate – – – ns 0.0005 ns – ns

Lee et al. (2007) (meta-analysis) 1168 Univariate 570 (49%) ns ns ns 0.000 – ns – 0.000
Costa et al. (2008) 49 Univariate 27 (55%) – – ns ns – ns ns ns
Elisei et al. (2008) 102 Univariate 38 (37%) ns ns 0.02 ns – ns ns 0.03

Wang et al. (2008a,b) 108 Univariate 54 (50%) ns – 0.02 0.02 – ns ns 0.04


Multivariate – – – 0.005 – – – –

Frasca et al. (2008) 323 Univariate 125 (39%) 0.005 – – 0.0001 – 0.0001 – 0.05
Multivariate – – – 0.0003 – 0.007 – ns

Lee et al. (2009) 64 Univariate 24 (38%) ns ns ns 0.001 – 0.003 – 0.001 (only T status)

Ito et al. (2009) 631 Univariate 242 (38%) – ns 0.05 ns – 0.005 ns ns


Multivariate ns (for tumors >1 cm) 0.033 (for ≥ 55 years) – – 0.0001 (for N1b) – –

Guan et al. (2009) 1032 Univariate 639 (62%) – ns ns 0.003 – 0.005 – 0.001
ns, not significant.
D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 15

Table 4
Association of BRAF with disease-free and overall survival.

Author Number of patients Disease-free survival (decrease) Overall survival (decrease)

Kim et al. (2006a,b) 203 ns –


Fugazzola et al. (2004) 260 ns –
Kebebew et al. (2007) 274 0.03 by multivariate logistic regression analysis –
Abubaker et al. (2008) 536 0.01 by Kaplan-Meier analysis –
Elisei et al. (2008) 102 0.03 by multivariate logistic regression analysis 0.015 by log-rank analysis
Ito et al. (2009) 631 ns –
Costa et al. (2008) 202 ns –

with overall survival. A higher rate of persistent disease, mortality, in diameter was referred to our department because of cyto-
or both in was reported in BRAF-positive PTC patients (P = 0.005). logic diagnosis of PTC. No enlarged lymph nodes were palpated
Multivariate logistic regression, that also included well-accepted or seen by sonography. The recommendation based on current
clinicopathological factors, found significant correlation with nega- guidelines (Cooper et al., 2009) would be to monitor carefully
tive outcomes only for BRAF status. Log-rank analysis confirmed the and, if no signs of distinct progression are observed, to perform
negative prognostic impact of BRAF mutation on overall survival, thyroidectomy after delivery. However, for better evaluation,
with deaths occurring practically exclusively in patients harbor- BRAF mutation was sought and confirmed to be present in the
ing BRAF-positive tumors. It must be stressed that patients in the cytologic material obtained by fine needle aspiration biopsy. Due
Elisei study were followed substantially longer than were patients to the presence of this molecular factor, the consulting surgeon
in studies negative for prognostic significance of BRAF mutation decided to perform surgery in the second trimester. In our opin-
(Fugazzola et al., 2004; Ito et al., 2009; Kim et al., 2006b). However, ion, the risk of overtreatment based on molecular diagnosis in
genetic background differences also may explain the discrepant this case was significant, since at least 20% of women of this age
results of the studies. presenting with PTC would bear BRAF-positive tumors, while the
A recent study on the molecular biology of recurrent PTC clinical data clearly show that the prognosis is excellent in these
(Henderson et al., 2009) is far less convincing than the Elisei et cases.
al. study (Elisei et al., 2008). The authors report that the BRAF On the other hand, the risk of undertreatment in the young
mutation was found in 77.8% of recurrences, which far exceeds BRAF-negative patients should be considered.
the frequency seen historically in primary PTCs. However, in the 2. The relevance of BRAF status to the indications for radioio-
majority of patients, these investigators did not compare the BRAF dine treatment seems still not well explained. On the one side,
status in the recurrent tumors versus in the primary tumors. there are plenty of data showing that BRAF mutation corelates
Additionally, they analyzed only lymph node recurrences, and with a poor response to radioiodine due to a lack of proper
the lower limit of their range of times-to-recurrence started at membrane sodium–iodine symporter expression (see further
4 months, an interval short enough to suggest that the disease discussion below). On the other side, there are proposals to
might be attributable to primary treatment failure rather than true apply more radioiodine in BRAF-positive patients due to the
recurrence. poorer prognosis. What we lack is evidence that more radioio-
dine will be helpful in these patients. Approaches to increase
3.2. Is it the time to apply BRAF as clinically relevant tumor radiosensitivity or to apply complementary pharmaco-
prognostic/predictive factor in PTC? logical therapy in these patients would appear to be more
reasonable.
The question, whether we are ready to implement the BRAF-
oncogene as a molecular prognostic factor in PTC is answered
positively by the majority of recent publications. However, the For the evaluation of the prognostic significance of BRAF in thy-
authors of this review would like to adopt a more cautious position roid cancer, a comparison with one of the most aggressive cancers,
and indicate what we believe is still necessary before BRAF might melanoma, may be of value. As already mentioned, BRAF mutations
be used as a clinically relevant factor of poor prognosis in DTC. Let are frequent in both cancers and are found at early disease stage—in
us list our major doubts: thyroid in micro-PTC and in skin, in the majority of benign nevi, the
precursors of melanomas. In benign nevi, the clonal activation of
1. As stated earlier, the frequency of BRAF mutation in PTC is MAPK by BRAF is followed by induction of senescence which pro-
high—45% on average, with values over 70–80% in some popula- vides a barrier against tumor progression (Wajapeyee et al., 2008).
tions. This high prevalence will mean that implementing BRAF Similarly, the expression of mutated BRAF in lung epithelium results
mutation as a factor of poor prognosis will shift many DTC in development of benign tumors that express senescence markers
patients, considered up to now as low risk patients, in the group and only rarely progress to malignant tumors unless functional p53
of patients who require more extensive therapy (Table 5). Our or p16 are absent (Dankort et al., 2007). BRAF-induced senescence
evaluation indicates that as many as 31% of all PTC patients has not been analyzed in thyroid tissues and can difficult due to lack
and 39% of those diagnosed with stage I–II disease, will face of benign counterpart to PTC. However, this hypothesis may explain
the risk of overtreatment if the decision will be based on the why BRAF is surprisingly frequent in indolent micro-PTCs—only the
BRAF-positivity of their tumors, among them at least of (1/4) of next molecular event(s) is/are necessary to accelerate its growth to
patients with micro-PTC, who constitute up to 44% of patients clinically evident disease. In this case, we should focus on identi-
with tumor ≤2 cm (Bonnet et al., 2009) and who were consid- fying events, secondary to BRAF mutation, which may be of better
ered until now as having an excellent prognosis. We think that, prognostic/predictive value then BRAF itself. According to Knauf
as of now, the evidence-based support for such consequences and Fagin (2009), in PTC, mechanisms for overcoming senescence
is very weak and the risk of overtreatment is significant. As an probably do not require loss of functional p53, which occurs late in
example, a 32-year-old pregnant woman (10th week of preg- progression to poorly differentiated thyroid carcinoma, nor rise in
nancy) with an incidentally detected hypoechoic nodule 8 mm p16.
16 D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28

Although molecular changes observed in other cancer cannot

General evaluation
be directly translated to PTC, we should also take into account
that in other types of cancer (e.g., colorectal, ovarian, lung) BRAF

Considerable

Acceptable

Acceptable

Acceptable
mutation is not related to poorer prognosis, on the contrary, it is
of the risk observed in cases with better outcome. In fact, the generally more
favorable outcome of BRAF-induced cancers may be the cause of
the generally good prognosis in PTC. This stresses once more the
need to search for molecular markers which shall stratify BRAF-
=A × (1 − B) × (1 − C)

positive PTC cases and help to specify patients with truly poor
In all PTC patients

prognosis.

3.3. Other molecular initiating events specific for PTC—do they


3%

3%
1%

3%

2%

9%
have prognostic potential?
Risk of undertreatment

3.3.1. RET/PTC rearrangements


=(1 − B) × (1 − C)

26%

6%
9%

5%

13%
In the subgroup

RET is a transmembrane tyrosine kinase receptor, the consitu-


tive activation of which by point mutation in parafollicular C cells
leads to medullary thyroid cancer. RET/PTC rearrangements, where
the 3 or tyrosine kinase domain of the RET gene is fused with the
5 domain of one of several constitutively expressed genes, cause
the constitutive activation of the RET gene, which otherwise is nor-
General evaluation

mally silent in follicular cells (Santoro et al., 2004). The fusion leaves
Theoretical consideration of the risk of over- and undertreatment if treatment of PTC is to be tailored according to the BRAF mutation status of the tumor.

intact the tyrosine domain (TK) of the RET receptor and enables
Not evaluable
Considerable

Considerable
d
Acceptable

Acceptable

the RET/PTC oncoprotein to activate the MAPK cascade (Knauf et


of the risk

al., 2003b). Several studies suggest that the oncogenic effects of


RET/PTC rearrangements require signaling along the MAPK path-
way in the presence of functional BRAF kinase (Knauf et al., 2003a;
Mitsutake et al., 2006).
In all PTC patients

RET/PTC rearrangements may arise from fusion with a wide


range of house-keeping genes (until now, 12 affected genes and
Therapeutic decisions based on

=A × C × B

circa 15 rearrangements have been described), and constitute an


event specific both for organ (thyroid) and histotype (PTC) (Fagin
1%

20%
4%

21%

10%

32%
the tumor BRAF status

and Mitsiades, 2008; Santoro et al., 2004). However, some data indi-
Risk of overtreatment

cate the possibility of RET/PTC rearrangements also in lymphocytic


In the subgroup

thyroiditis (Rhoden et al., 2006).


Transfection of the RET/PTC1 gene in normal rat thyroid cells
resulted in loss of differentiation and of TSH growth dependency.
=C × B

However, the cells were totally transformed only after transfection


9%

39%
37%

39%

39%

with RET/PTC and mutated RAS genes, suggesting that simultaneous


activation of several genes was necessary for tumor progression.
5 Years DFSa

This observation, together with the frequent occurrence of RET/PTC


CSS, cancer-specific survival; DFS, disease-free survival; PTC, papillary thyroid cancer.

rearrangements in papillary microcarcinoma, suggests that such


70%

90%
85%

70%

rearrangements are an early event in thyroid carcinogenesis (Jhiang


∼90%


(C)

et al., 1996; Viglietto et al., 1995).


A wide range of RET/PTC prevalence in PTC, ranging from 3% to
5 Years CSSa

85% has been reported (Kondo et al., 2006; Nikiforov, 2002; Tallini
and Asa, 2001). In general, RET/PTC prevalence is higher in young
100%

100%
98%

99%

95%
70%

patients and in radiation-induced PTC, and decreases with patients’


Data extracted from the meta-analysis of Lee et al. (2007).

age (Collins et al., 2002; Fenton et al., 2000). However, apart from
study population variations, the wide range of RET/PTC prevalence
Younger than 20 years, usually in >15 years patients.
13%

43%

56%

also reflects the different procedures used to identify RET/PTC rear-


% of patients with
BRAF mutationsb

rangements (Zhu et al., 2006). RET rearrangement is frequently


Because of poorer prognosis in this subgroup.

found in micro-PTC, an observation that is usually interpreted as


Data estimated from Jonklaas et al. (2006).

the evidence in favor of an initiating role in PTC. However, doubts


about this interpretation have been raised by the recent reports on
(B)

RET/PTC multiclonality (Unger et al., 2004), which may suggest that


Young patients (≤20 years of age)

at least in some PTCs, the recombinations may have arisen later in


% of patients

tumor progression.
with PTCa

The majority of RET/PTC studies did not find any correlation


between the rearrangements and other clinicopathological factors
10%

40%
20%

53%

26%
4%
(A)

related to PTC course or outcome. In few of the studies, though,


Adult patients

an increased frequency of lymph node metastases was noticed in


All stages

patients with RET/PTC rearrangements (Adeniran et al., 2006). Ini-


tially, this observation was interpreted as evidence for correlation
I + IIc

I + II
Table 5

Stage

IV
III

with poorer prognosis. Later on, when the more frequent occur-
II
I

rence of RET/PTC in younger of PTC patients has been known, the


D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 17

higher occurrence of lymph node metastases was correlated with a variety of membrane receptors, mainly belonging to the tyrosine
younger age rather than with RET/PTC rearrangements (Collins et kinase receptor family, rarely G proteins. Non-membrane coupled
al., 2002). tyrosine kinases can also to activate RAS, with the down-stream
activation of different effector pathways.
3.3.2. NTRK rearrangements The RAS protein is encoded by three genes, H-, N- and K-RAS
NTRK1 is another membrane tyrosine kinase receptor which reg- and each of them may underlie activating point mutations, occur-
ulates growth, differentiation and apoptosis in the peripheral and ring most frequently at codon 12 or 61. Such mutations lead to the
central nervous system, with nerve growth factor acting as its lig- change in conformation of the protein so that it remains bound
and. Similarly to RET, NTRK-3 is activated by rearrangements with to GTP, i.e., constitutively active (while in normal conditions GTP-
genes constitutively expressed in thyroid follicular cells. Activation ase activity of RAS leads to RAS deactivation). In some cancers, like
of the oncoproteins TRK1-3 is ensured by autophosphorylation of colon cancer, RAS mutation is regarded as the first oncogenic event
the tyrosine kinase domain (Pierotti et al., 1996). NTRK rearrange- necessary but not sufficient for the cell to cross over the benign ade-
ments are rare and are found in <10% of PTCs (Brzezianska et al., noma/cancer barrier (Takayama et al., 2006) as RAS mutations are
2007; Roque et al., 2001). Their rarity precludes the studies on their found with the same frequency in colonic adenomas as in colonic
prognostic roles. cancers. In some other types of cancer RAS mutations occur at later
stages of carcinogenesis.
3.3.3. Gene rearrangements are frequent in thyroid cancers In thyroid gland, RAS mutations are found already in follicu-
Until recently, gene rearrangements were regarded as charac- lar thyroid adenomas (FAs) and the transition from the follicular
teristic of hematological malignancies. Among solid tumors, they adenoma to follicular cancer requires the next molecular step
were found only in sarcomas and thyroid carcinomas. This picture (although, unlike in colon cancer, adenomas are not regarded as
is now being changed by the detection of very small chromoso- precancerous stage for FTC). RAS mutations were described in some
mal deletions/amplifications in other solid tumors, e.g., prostate 19% of FAs and in up to about 50% of FTC (Esapa et al., 1999; Shi et
cancer (Carver et al., 2009; Tomlins et al., 2009). Nevertheless, al., 1991; Suarez et al., 1990), however, in more recent studies RAS
predisposition of thyroid cells to undergo gene rearrangement is mutations are evaluated with an overall frequency of about 20%
not completely understood (Teixeira, 2006; Wang et al., 2008b). (Garcia-Rostan et al., 2003; Vasko et al., 2003).
The juxtaposing of the loci participating in RET/PTC and NTRK1 Importantly, RAS mutations are not thyroid cancer histotype-
rearrangements in interphase nuclei of thyroid follicular cells has specific and are also found in PTC, with the prevalence ranging from
been reported (Gandhi et al., 2006; Nikiforova et al., 2000; Roccato 0% to 15% (Abrosimov et al., 2007b; Adeniran et al., 2006; Hou et al.,
et al., 2005). Interestingly, these rearrangements are more fre- 2007; Vasko et al., 2003). These mutations are rarest in the classical
quently found in young patients (Collins et al., 2002; Fenton et variant of PTC, and more frequent in the follicular variant, where
al., 2000; Wang et al., 2008a). Of note, though, the decisive fac- their prevalence may reach even 20–50% (Goutas et al., 2008; Zhu
tor(s) explaining the good prognosis of PTC in young patients, et al., 2003). Some studies indicate that N2-RAS mutations are par-
remains unclear: is it the biology of the tumor (which is caused ticularly frequent (Abubaker et al., 2008; Hou et al., 2007; Wang
by a gene rearrangement, not BRAF mutation) or of the host et al., 2007), other focus on K-RAS (Goutas et al., 2008) which are
(different immunity in young patients) or of both (Jarzab et al., seen in FTC with lower frequency (Fagin and Mitsiades, 2008). Their
2005a). frequency in anaplastic thyroid cancer is estimated as to be about
20–25% (Smallridge et al., 2009).
3.4. Is there some multiclonality in papillary thyroid cancer? RAS mutation can promote thyroid tumorigenesis through
the RAF–MEK–ERK1/2 pathway or through its interaction with
Many data indicate that, despite the general rule of non-overlap PI3K/AKT pathway (Xing, 2005). Garcia-Rostan et al. (2003)
between the main PTC-inducing mutations, which are regarded as reported the association of RAS mutations with aggressive thyroid
mutually exclusive (Fagin, 2004), there is a substantial degree of cancer phenotypes and poor prognosis: 55% (11/20) of patients
multiclonality in this type of thyroid cancer. Both BRAF mutations with RAS-positive well-differentiated thyroid cancers died in com-
and RET/PTC rearrangements can be found as non-clonal changes parison to 15% (9/58) RAS-negative ones (P = 0.016). There are no
and different types of RET/PTC rearrangement may be found in dif- other reports on this issue, however, the observation that RAS muta-
ferent tumor foci from the same patient (Aherne et al., 2008; Oler et tions are rather frequent in poorly differentiated thyroid cancers
al., 2005; Sugg et al., 1998; Unger et al., 2004; Vasko et al., 2005; Zhu (40–55%) supports this notion (Fagin and Mitsiades, 2008).
et al., 2006). Concomitant presence of BRAF mutation and RET/PTC
rearrangement in one tumor is possible and this is also true for 3.6. Follicular thyroid cancer-specific gene mutations
other mutations (Nikiforova et al., 2004; Smyth et al., 2005; Wang
et al., 2008b). This fact is of a great potential impact on the power 3.6.1. PAX8/PPAR rearrangements
of these molecular changes as prognostic or predictive factors. This type of rearrangement results from the translocation
between the PAX8 (paired box gene 8) gene, which encodes a paired
3.5. Gene mutations important for development of both papillary domain transcriptional factor, and the peroxisome proliferator-
and follicular thyroid cancer activated receptor (PPAR) gene. The resulting fusion protein (PPFP;
PAX8/PPAR␥ fusion protein) encodes a nearly full-length PPAR␥,
3.5.1. RAS mutations the expression of which is under the transcriptional regulation of
The RAS–RAF–MEK–ERK1/2 pathway is hyperactivated in 30% the PAX8 promoter (Kroll et al., 2000). The functional consequences
of human cancers (Balmanno and Cook, 2009). RAS is a membrane- of this PAX8/PPAR rearrangement are yet not fully understood.
associated, small protein with guanosine triphosphate (GTP) Some studies have shown negative effect of PAX8/PPAR rearrange-
binding ability, involved in proliferation, differentiation and cell ment on the function of wild-type PPAR (Gregory et al., 2004),
survival. Translocation of RAS to the cytoplasmatic membrane is however other have questioned this negative effect and showed
an important step in the protein’s activation. Farnesylation of RAS the up-regulation of many genes which are transcriptional tar-
is the first obligatory step in a series of post-translational modifica- gets of PPAR (Giordano et al., 2005). Another possible oncogenic
tions leading to membrane association, which, in turn, determines mechanism of PAX8/PPAR rearrangement is deregulation of PAX8
the switch from an inactive to an active form. RAS is activated by function, which is critical for thyroid cell differentiation (Reddi
18 D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28

et al., 2007). In FTC, PAX8/PPAR rearrangement is expressed in Mitsiades, 2008). The Wingless (Wnt) family of secreted glyco-
some 35–47% of cases (Cheung et al., 2003; Fagin and Mitsiades, proteins controls early developmental processes including cellular
2008; Sahin et al., 2005). PAX8/PPAR rearrangement has also been migration, proliferation and differentiation. Non-canonical Wnt
shown in follicular adenoma with a varying frequency (Cheung et signaling is associated with tumorigenesis and Wnt5a has been
al., 2003; Marques et al., 2004), summed up to 13% (Sahin et al., reported to be increased in the majority of PTCs but in only some
2005), thus may be regarded as an early event in follicular tumori- FTCs (Kremenevskaja et al., 2005). Mutations of ␤-catenin gene
genesis, requiring additional mutations to induce FTC (Cheung et (CTNNB1) are most frequent in poorly differentiated and anaplastic
al., 2003; Gregory et al., 2004). Again, RAS and PAX8/PPAR are cancer (Garcia-Rostan et al., 1999; Garcia-Rostan et al., 2001).
regarded mutually exclusive (Nikiforova et al., 2003b). There are
no very convincing data on PAX8/PPAR rearrangement prognos- 3.7.4. TP53 mutations
tic significance (Sahin et al., 2005). The rearrrangement does not TP53 encodes a multifunctional nuclear protein which is impor-
appear to be expressed in classical PTC, while in follicular variant tant for cell cycle arrest at DNA damage, senescence or apoptosis
of PTC some authors describe it frequent occurrence (Castro et al., and is one of the best known tumor suppressor genes. Contrary
2006), while others (Sahin et al., 2005) negate this possibility. to what is seen in many other cancers, TP53 loss of function
mutations occur late in thyroid tumorigenesis and are practically
3.6.2. RAS homolog 1 (ARH1) silencing absent in DTC (0–9%), while the prevalence of these mutations
As already mentioned, the mechanisms of transition from a rises as the tumor grade worsens, reaching 17–38% in poorly dif-
benign to a malignant tumor, from FA to FTC, are not well known. ferentiated and 55–88% (Fagin and Mitsiades, 2008; Smallridge
High frequency of loss of heterozygosity in imprinted genomic et al., 2009) in anaplastic cancers. At this moment, the features
regions has been suggested to contribute to this transition (Sarquis of poor differentiation are easily detected histologically, thus, the
et al., 2006). One of the proposed mechanisms is the reduction of the immunohistochemical or molecular investigation of TP53, docu-
number of ARH1 copies. Weber et al. (2005a) showed that relatively mented widely in other tumors, has not gained clinical significance
few FAs but the majority of FTCs, including minimally invasive FTCs, in thyroid cancer.
show marked ARH1 mRNA underexpression, due to deletion of the
non-imprinted second allele in conjunction with hypermethylation 4. Prognostic aspects of other gene expression changes in
of the genomically imprinted allele. papillary and follicular thyroid cancer

Over the last two decades many gene expression changes have
3.7. Mutations of other genes and their importance for thyroid
been described in papillary and follicular thyroid cancers and
cancer and its outcome
related to the stage of the disease or its outcome. The clear distinc-
tion between the changes characteristic for papillary and follicular
3.7.1. PIK3CA mutations and amplification
cancers is however not possible for many reasons: First, due to
PIK3CA mutation is not a common mechanism in the activation
the much higher frequency of PTC, most of the genes described
of PI3K/AKT in thyroid carcinoma while amplification of this gene
below were either not examined in FTC or investigated in small
is more frequent. In PTC, rare mutations and much more frequent
groups of FTC cases, not allowing for specific distinction. Second,
amplifications, comprise alltogether 15–53% of tumors (Abubaker
many of those genes in which such comparison was performed,
et al., 2008; Hou et al., 2007). PIK3CA amplication is observed also in
showed only minor differences between histotypes because they
FTC and even in FAs (Hou et al., 2007; Wu et al., 2005); however, it
were late consequences of the neoplastic transformation of the fol-
is the most frequent in anaplastic thyroid cancer. Its down-stream
licular thyroid cell. Thus, it is rationale to look for their prognostic
target, AKT, is phosphorylated in at least half of PTC, but indepen-
significance independently from the histotype or variant of thyroid
dently of PIK3CA mutation or amplification (Abubaker et al., 2008).
cancer. In the given below brief list of genes, expression changes of
which has putative or proved prognostic relevance, we inform on
3.7.2. PTEN mutations histotype-specific changes only then, when they were investigated
PTEN (phosphatase and tensin homolog) is a dual-specific with sufficient power.
phosphatase, a negative regulator of the AKT/PI3K pathway
by dephosphorylation of phosphatytylinositol-3,4,5-triphosphate 4.1. Single genes expression changes
(PIP3), which can also influence the MAPK pathway. PTEN is a tumor
suppressor gene and its germline loss of function mutations pre- 4.1.1. MET proto-oncogene
dispose to multiple tumors including FTC (Yeh et al., 1999). Single The MET proto-oncogene encodes a membrane tyrosine kinase
cases of loss of function somatic mutation of PTEN were described receptor for hepatocyte growth factor (HGF). HGF is a potent mito-
in FTC (7%), somewhat more frequently in ATC (12–50%) and rarely gen for epithelial cells and promotes cell motility and invasion.
in PTC (2%) (Frisk et al., 2002; Garcia-Rostan et al., 2005; Gimm et About 50% of cases of PTC are characterized by MET overexpression
al., 2000; Hou et al., 2007; Smallridge et al., 2009). A novel rear- (Di Renzo et al., 1992), which is believed to be a sign of more aggres-
rangement involving PTEN and the H4 gene, has been described as sive disease (Mineo et al., 2004; Ramirez et al., 2000). BRAF-positive
the non-clonal event in 14/18 PTCs investigated (Puxeddu et al., tumors were associated wit MET overexpression in aneuploid PTCs
2005) and as a feature present in 4.8% of a larger series of PTCs in (Rodrigues et al., 2007).
Chinese patients (Wang et al., 2008b). All these rearrangements are
paracentric inversions of chromosome 10q, the same chromosome 4.1.2. Epithelial growth factor receptor (EGFR)
which is responsible for RET rearrangements. Among receptor tyrosine kinases EGFR belongs to the best
known. EGFR (HER1) protein levels are increased in PTCs in com-
3.7.3. ˇ-Catenin mutations parison to normal thyroid tissue and the high expression of EGFR
The ␤-catenin protein, when bound to E-cadherin, mediates has been suggested to be associated with worse outcome of PTC
cell skeleton/adhesion interactions, and when non-sequestered after thyroidectomy (Akslen and Varhaug, 1995), an observation
and non-degraded, plays a role in gene transcription regulation not repeated in other studies (Ruan et al., 2008). Another early
of growth-promoting genes. It constitutes an element of Wnt sig- study indicated on HER2 as the predictor of metastasis in PTC
naling pathway and is in fact a Wnt nuclear effector (Fagin and (Kremser et al., 2003). The immunohistochemical evaluation of
D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 19

EGFRvIII showed its expression in the majority of PTCs (Omidfar 4.1.6. Cyclin D1 and other cell cycle regulators
et al., 2009). In the study of Wiseman et al. (2008a), which investi- Association of overexpression of cyclin D1, one of the target
gated expression of HER1–4 genes, overexpression of at least one genes of Wnt signaling pathway, with more advanced PTC, has
of them was observed in 76% of DTC and only HER3 was positively been adressed for a long time (Ferenc et al., 2005; Ito et al., 2005).
correlated with DTC stage, while HER4 was inversely correlated Although cyclin D1 and other cell cycle regulators, such as cyclin E,
with T stage. In this context it is important that HER3 (ERBB3) is p16, p21 were overexpressed in thyroid cancer, only p16 expres-
characteristic for BRAF-induced PTCs (Giordano et al., 2005). sion correlated significantly with extrathyroidal tumor extension
Until recently, it was assumed that activating mutations of EGFR and the presence of lymph node metastases in the study of Melck
do not play any role in the molecular biology of DTC (Omidfar et et al. (2007). Also, in the study of Pesutic-Pisac et al. (2008) loss
al., 2009). However, the recent finding of Masago et al. (Masago et of p27 was more essential than cyclin D1 overexpression, while
al., 2009) indicated that EGFR mutations commonly seen in pul- Khoo et al. (2002) suggested the contribution of both markers. The
monary carcinoma may be observed in a subset of papillary or study of Hoos et al. (2002) was negative for correlation of cyclin
poorly differentiated thyroid cancers which may make tumor cells D1 expression with indices of prognostic relevance in oncocytic
oncogene-addicted. thyroid cancers.

4.1.7. Mucin 1
4.1.3. Mitogen-inducible gene-6 (MIG-6) Mucin 1 (MUC1) is a glycoprotein important for cell adhe-
MIG-6 is an immediate early response gene, expression of which sion, overexpressed in about 25% of PTC (Wreesmann et al., 2004).
is induced by cellular stress, hormones or growth factors in many Preclinical studies support the role of MUC1 in promoting an
cells. It exerts a negative effect on EGFR signaling that is mediated aggressive phenotype of PTC (Patel et al., 2005). MUC1 overexpres-
by the MAPK cascade (Ferby et al., 2006). Because MIG-6 expres- sion is especially frequent in the tall-cell variant of PTC, known
sion is induced by MEK–ERK activation, such expression may be for its poor prognosis (Ghossein and LiVolsi, 2008). Identified as
regarded as a negative feedback loop signal in normal cells, lost an independent prognostic marker by genome-wide gene expres-
in cancer. In PTC, MIG-6 directly correlated with EGFR expression sion profiling of PTC, MUC1 overexpression was described to be
and its higher mRNA level was associated with better overall sur- significantly associated with the treatment outcome (Wreesmann
vival in a population of 106 PTC patients followed for six years. et al., 2004), independent of extrathyroidal extension or of the PTC
Interestingly, MIG-6 expression was independently predictive of variant. However, other reports question the correlation of MUC1
disease-free survival in BRAF-positive patients (Ferby et al., 2006). overexpression with PTC prognosis (Abrosimov et al., 2007a; Min
et al., 2008). The significance of MUC1 overexpression for cancer
outcome may be more essential if we consider MUC1 contribution
4.1.4. Vascular epithelial growth factor (VEGF) and VEGF receptor to the survival network, causing resistance to genotoxic agents in
VEGF overexpression is a characteristic feature of malignant thyroid cancer (Siragusa et al., 2007).
tumors, seen in thyroid cancer as in other cancers (Wiseman et al.,
2008b) and may be interpreted as marker of tumor hypoxia. How- 4.1.8. S100A4 calcium-binding protein
ever, not in all cancers it is correlated with HIF-1alpha expression. This small calcium-binding protein also called calvasculin has
Thyroid cancer belongs to those tumors in which hypoxia-inducible been reported to promote metastasis and was indicated as a molec-
factor-1␣ is up-regulated at the same time that VEGF is overex- ular factor of poor prognosis in PTC (Zou et al., 2004, 2005). In
pressed (Jubb et al., 2004). The correlation of VEGF expression with papillary microcarcinomas, the expression of S100A4 correlated
the expression of other angiogenic factors is strong (Tanaka et al., more strongly with lymph node metastases than did any other
2002). Quantitative evaluation of VEGF expression in PTC showed molecular factor investigated and was more significant for asso-
the association of this phenomenon with metastatic PTC (Klein et ciation with stage than BRAF status (Min et al., 2008).
al., 2001). VEGF expression was closely correlated with tumor size,
extrathyroidal invasion and BRAF presence (Jo et al., 2006), while 4.1.9. Osteopontin and CD44v6
Tian et al. (2008) observed the correlation with lymph node metas- Osteopontin (SPP1) is a cytokine regulating cell trafficking
tases both for VEGF and metalloproteinase MMP2. In the study of Jo within the immune system, binding a splice variant of CD44,
et al (Jo et al., 2006), VEGF was the only molecular marker associated that is overexpressed in PTC and contributes to the mitogenesis
with tumor size, extrathyroidal invasion and cancer stage by uni- (Figge et al., 1994), survival and motility of PTC thyrocytes and their
variate analysis and the association with extrathyroidal invasion invasion potential (Guarino et al., 2005). In PTC, the gene expres-
was confirmed in multivariate analysis, while VEGFR expression sion level of SPP1 was associated with metastasis but not with the
was only weakly associated with PTC recurrence in young patients BRAF status of the tumor (Oler et al., 2008).
(Fenton et al., 2000).
4.1.10. Immunity-related genes
Activation of the MAPK pathway induces up-regulation of
4.1.5. E-cadherin and catenins chemokines and their receptors on tumor cells, relevant for their
E-cadherin (CDH1) is a calcium-dependent transmembrane gly- sustained proliferation and cell motility. Tumors use molecules of
coprotein that functions as a cell-cell adhesion molecule. It has an the innate immune system for their growth, survival and metasta-
intracellular domain that complexes with catenin proteins for regu- sis (Melillo et al., 2005). Expression of toll-like receptor 3 (TLR3) is
lation of further Wnt-mediated signal transduction to the nucleus. characteristic for PTC but not FTC (McCall et al., 2007). The basally
CDH1 expression is reduced in thyroid carcinomas and has been expressed receptors are functional and able to activate the NF␬B
associated with poor prognosis, not only in PTC (von Rhein et al., pathway and to increase the expression of interferone IFN␤ and
1997) but also in FTC (Brecelj et al., 2005) as well as with PTC-to- CXCL10 chemokine. Interestingly, TLR5 overexpression is coordi-
ATC transition (Wiseman et al., 2007). Dislocalization of catenins in nated with WNT5a expression.
thyroid cancer may be secondary to loss of E-cadherin (Bohm et al., Chemokine receptors CXCR2–CXCR4 and CCR7 are expressed on
2000; Rocha et al., 2003). However, the study of Kapran et al. (2002) PTC cells (Melillo et al., 2005; Sancho et al., 2006; Wagner et al.,
did not show any association of E-cadherin or catenins expression 2008) and CCR7 was reported to be associated with pathologic fac-
with PTC recurrence. tors of PTC aggressiveness such as extrathyroidal tumor extension,
20 D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28

angiolymphatic invasion and presence of lymph node metastases more sensitive. Unfortunately, this gene signature has not yet been
(Wagner et al., 2008). Chemokine CXCL14 was described to be asso- verified by another group.
ciated with metastatic PTC (Oler et al., 2008).
Interestingly, expression of HLA-DR and DQ molecules is 4.2.2. Metastasis-associated genes in thyroid cancer
inversely associated with PTC recurrence risk (Jo et al., 2008). Although the survival rate in well-differentiated thyroid cancer
is excellent, that is not the case in the 5–10% of patients with dis-
4.2. Prognosis-related gene signatures in thyroid cancer tant metastases. Thus, identification of a metastasis-related gene
signature is desirable. One of the first approaches was carried out
4.2.1. Gene expression profiling and prognosis of thyroid cancer by Zou et al. (2004), who established a thyroid cancer cell line with
Several DNA microarray studies show distinct differences in the high metastatic capacity and specified genes characteristic of this
molecular profiles of not only papillary versus follicular cancer, but line, among them MET, ezrin, integrin, motility-related protein-
also between subtypes of papillary cancer (Aldred et al., 2004; Baris 1, cadherin P, NEDD5 and S100A4. In their quantitative real-time
et al., 2005; Chevillard et al., 2004; Finley et al., 2004a,b; Giordano PCR (QPCR) validation study, mentioned already above, they con-
et al., 2005; Oler et al., 2008). It is to be stressed that some gene firmed S100A4 to be associated with poor thyroid cancer prognosis.
ontology groups are characteristic of PTC, mainly cell adhesion- Two other genes were listed by Oler et al. (2008) as metastasis-
related genes involved in cancer invasion and metastasis, or the associated; CXCL14 and cystatin M. However, the definition of
gene signature related to the immune response to this DTC his- metastatic PTC was very poor in this study. Some comparisons
totype (Huang et al., 2001; Jarzab et al., 2005b). While the gene were also performed by Rodrigues et al. (2007) who reported low
expression profile of PTC differs from that of normal thyroid tissue expression of many genes in M1 aneuploid PTCs.
by a huge number of overexpressed genes, in FTC, down-regulated
genes far outweigh up-regulated genes (Aldred et al., 2003). In the 4.2.3. Epithelial-to-mesenchymal transition in PTC
recent study by Oler et al. (2008) based on SAGE (serial analysis of In the search for key molecular prognostic factors in thyroid
gene expression) and followed by QPCR (quantitative polymerase cancer, the mechanisms of tumor invasion must be considered.
chain reaction), there were 548 differentialy expressed genes when The process of epithelial-to-mesenchymal transition (EMT) has
PTC was compared with FTC and FA. Of the 548, 131 genes had a been implicated in the conversion of early-stage tumors to inva-
more than 10-fold increase in expression in PTC, and 61 genes, an sive malignancy and has been extensively studied in invasive PTC
over 15-fold increase. (Vasko et al., 2007). After loss of cell-to-cell contacts and remodel-
Despite the very characteristic morphologic phenotype of ing of the cytoskeleton, tumor cells are more prone to migration and
PTC, surprisingly few single genes were described as unique for this process is hallmarked by loss of cadherin-E expression. Inte-
this histotype. CITED1 (Cbp/p300-interacting transactivator, with grin pathway, Notch, MET, TGF-beta, NF␬B, PI3K and p21-activated
Glu/Asp-rich carboxy-terminal domain, 1), encoding a transcrip- kinase belong to EMT-regulating proteins, also fibronectin 1 is
tional transactivator nuclear protein, is one of the few identified considered as EMT related. In the study of Vasko et al. (2007),
until now (Prasad et al., 2004). Other examples are nicotinamide- microarray analysis of the central part and of the invasive front
N-methyl-transferase (NNMT), encoding a liver enzyme involved of selected PTC tumors has been performed, showing in the lat-
in the biotransformation of many drugs and xenobiotics (Xu et al., ter the overexpression of TGF-beta, NF␬B and integrin pathway as
2003), or chemokine (C–X–C motif) ligand 14 (CXCL14), cystatin well as of small G protein regulators and CDC42. Overexpression of
E/M (CST6) and DHRS3 (dehydrogenase/reductase (SDR family) vimentin, the hallmark of EMT, was associated with tumor invasion
member 3), a protein involved in retinol storage, all three found and nodal metastasis and was correlated with RUNX2 expres-
to be characteristic of PTC by Oler et al. (2008). It seems that rather sion, as judged by immunohistochemistry. Interestingly, there was
than single genes, gene groups (gene clusters or gene signatures) no correlation of expression of these invasion-related proteins
are characteristic of PTC, among them, the already-mentioned gene with tumor genotype, e.g., BRAF status. However, in the study of
classes related to adhesion/motility or invasion (Fig. 2). On the other Watanabe et al. (2009) silencing of BRAF reduced the increased
hand, many genes have similarly altered expression in both PTC expression of vimentin. For PTC cell invasion, enhanced AKT activ-
and FTC, which helps to define these changes as common conse- ity is especially important, but there was no correlation of AKT with
quences of different neoplastic transformation pathways (Polanski BRAF presence in PTC (Vasko et al., 2007). According to Wang et al.
et al., 2007). (2007), BRAF mutation and AKT pathway activation are mutually
In FTC, according to Aldred et al. (2003), three of the most down- exclusive in PTC.
regulated genes localized to previously reported regions of loss of
heterozygosity (LOH) were caveolin-1 and -2 and bone morpho- 4.3. Genetic germline background and the prognosis of papillary
genetic protein 3b (GDF10). Takano et al. (2004) identified the trefoil and follicular thyroid cancer
factor 3 gene as down-regulated in FTC, but this gene should not
be regarded as FTC-specific, as it is down-regulated in PTC as well Some 5–10% of all cancers developing in follicular cells occur on
(Hamada et al., 2005). the background of familial predisposition, referred as familial non-
Despite the publication of numerous thyroid cancer gene medullary thyroid carcinoma (FNMTC), with autosomal dominant
expression profile analyses (see the meta-analysis of Griffith et mode of inheritance and penetrance increasing with age (Cavaco
al., 2006), only a few studies addressed prognosis-related gene et al., 2008; Handkiewcz-Junak et al., 2006; Malchoff and Malchoff,
signatures. Finley et al. (2004b) were the first to assess whether 2006). This clinical entity develops most often (in over 85% of
gene expression profiling was capable of differentiating aggressive cases) as familial PTC and differs from sporadic cancer by earlier
thyroid carcinomas from non-aggressive ones. They characterized age of onset, more frequent multifocal disease, higher degree of
their group of 29 DTC tissue samples according to the patients’ aggressiveness and higher recurrence rate. Not all involved family
AMES criteria and obtained a list of 339 differentially expressed members suffer from thyroid cancer, some develop multinodular
genes. By this criterion, all of an additional 6 samples for which goiter or follicular adenomas. The diseases are genetically heteroge-
AMES status was unknown were correctly clustered, while there nous, with some genes identified in single families but with only
were 6/29 misclassified samples in the test group, yielding a sen- a few general susceptibility genes described (Canzian et al., 1998;
sitivity of 85.7% and a specificity of 80.9% for high-risk tumors. Gudmundsson et al., 2009; Jazdzewski et al., 2008; McKay et al.,
The same procedure done for extrathyroidal extension was even 2001; Ngan et al., 2009). Somatic BRAF or RAS mutations are found in
D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 21

Fig. 2. A large cluster of cell adhesion-related genes are homogenously up-regulated in PTCs. By their expression nearly all PTCs may be separated from follicular tumors and
benign tissues. Hierarchical cluster analysis based on 100 PTC and 99 FTC, FA and non-neoplastic thyroid samples from own material (49 PTC tumors) and those published
by Giordano et al. (2005) and Weber et al. (2005b). The intensity of the color refers to the intensity of gene expression; red refers to the up-regulation, and green refers to
the downregulation. Color coding of tissues: blue, PTC; red, FTC; green, benign tumors.

familial PTCs with similar frequency as in sporadic tumors (Cavaco probesets with P < 0.01 were specified, with a similar distance of
et al., 2008). In this aspect it is important to add that there are no both of these mutant types from RAS mutants, the number of dif-
data suggesting that any particular genetic background facilitates ferent genes being on average approximately 20 times as many as
the BRAF mutation what could be suspected considering the pres- obtained for permuted data sets. By further selection of genes for
ence of a unique BRAF mutation present in a large proportion of PTC those that yielded P < 0.01 as well as a fold-change <0.5 or >2 for any
cases. However, environmental influences are also conceivable, as group compared to any other, the BRAF mutants were characterized
suggested by recent Chinese data showing that BRAF-positive PTCs by 82 probesets (31 up-, 51 down-expressed). These probesets had
are significantly more frequent in areas of extremely high iodine a very strong significance, verified, as required in genomic studies,
intake (Guan et al., 2009) and by comparison of different Sicilian by the False Discovery Ratio (FDR), which was 0.5 (it is well accepted
areas (Frasca et al., 2008). in genomic studies that significant genes should show FDR under
5–10%). Many genes were down-regulated in RAS mutants, while
5. Final considerations on prognostic impact of genetic many increases were common between BRAF and RET/PTC mutant
factors in papillary and follicular thyroid cancer tumors. The most differentially expressed gene in the BRAF mutants
was TM7SF4. Its product is a dendritic cell protein, likely related to
5.1. Relation between BRAF status and gene expression profile the immune response in these tumors.
Although the other studies published were less powerful
5.1.1. Is BRAF-positive DTC a distinct molecular subtype? (Frattini et al., 2004; Musholt et al., 2006), they did not contra-
The mutational status of PTC correlates well with its gene dict the basic message of the Giordano study: the gene expression
expression profile, allowing bioinformatical analysis of genome- profile of BRAF-positive PTC is very distinct and easily recognizable.
wide data to clearly differentiate BRAF, RAS or RET/PTC mutants, Oler et al. (2008) described that the expression of a newly discov-
even if an unsupervised approach, Principal Component Analysis, ered chemokine, CXCL14, and of cystatin M, a secreted inhibitor
is applied. The most important paper on this issue, published by of lysosomal cysteine proteases, was not only markedly elevated
Giordano et al. (2005), showed distinctly that PTC mutational sta- in PTC, but also significantly higher in BRAF-positive versus BRAF-
tus was even more strongly correlated with gene expression profile negative tumors. Both genes were simultaneously associated with
than with morphology as evaluated by subdivision into PTC vari- metastatic PTC. Also, expression of such PTC-characteristic genes
ants. When comparing RET/PTC mutants versus BRAF mutants, 3891 as fibronectin 1, CITED1 and vimentin were positively correlated
22 D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28

with the expression of BRAFV600E (Watanabe et al., 2009). The (1.01% versus 0.1%) (Nakayama et al., 2007); importantly, both
conditional expression of BRAFV600E and RET/PTC showed that the these Ki-67 values are rather low, which is a known feature of PTC
more invasive potential of BRAF-mutated tumors may be at least (Saltman et al., 2006). The increased GLUT-1 expression in BRAF-
partially explained by more distinct overexpression of metallopro- positive tumors may be interpreted as the sign of such tumors’
teinases MMP3, MMP9 and MMP13 (Mesa et al., 2006). higher aggressivity. A multivariate analysis concluded that BRAF-
All these data, especially regarding genome-wide differences, positive tumors have higher VEGF expression (Jo et al., 2006).
substantiate the hypothesis that BRAF-induced PTC is in fact a dis- Recent reports indicate that prohibitin, a multifunctional protein,
tinct molecular subtype of PTC, with a different, more aggressive is overexpressed only in PTCs bearing a BRAF mutation. The most
disease course and a poorer outcome, which demands a subtype- plausible mechanism for prohibitin overexpression is the increased
specific treatment. However, prospective data still are necessary to gene promoter activity by the BRAF oncoprotein (Franzoni et al.,
prove this theory. 2009).
Nevertheless, recent publications indicate that some molecu-
5.1.2. BRAF-mutated PTCs exhibit fewer thyroid-specific gene lar markers correlate more closely with clinicopathological factors
expression features of poor prognosis than does BRAF-oncogene presence itself. In
Many authors have confirmed the lower expression of thyroid- contrast to BRAF-positivity, S100A4 and cyclin D1 overexpression
specific genes in BRAF-positive PTCs: the gene expression profiling predicted lymph node metastasis in a series of nearly 200 PTC
study by Giordano et al. (2005) noted a down-expression of TPO microcarcinomas (Min et al., 2008). The S100A4 gene was also
and the study of Durante et al. (2007) confirmed the earlier obser- investigated in a more detailed study and shown to be significantly
vations of the more intensive loss of gene expression of NIS (−82%), associated with metastases (Zou et al., 2005).
AIT-B (−86%), TPO (−90%), and TG (−46%) in BRAF-mutated PTCs.
When BRAF-positive and BRAF-negative PTCs are compared, the
level of NIS gene expression is at least 5 times lower in the former 5.2. Major questions in the therapeutic strategy in DTC which can
(Durante et al., 2007). Not only is there a BRAF-dependent down- be solved by molecular diagnosis
regulation of NIS expression on the transcriptional level, but also,
the trafficking of the NIS protein to the cell membrane is impaired in In fact, currently, the major clinically relevant problems we face
BRAF-positive PTCs (Riesco-Eizaguirre et al., 2006). Moreover, sup- in DTC are:
pression of the BRAF/MEK/MAPK pathway restores expression of
iodide-metabolizing genes (Liu et al., 2007a). Unlike the case with
NIS or TPO genes, the expression of TSH-R and PAX8 seems not to 1. How to differentiate between indolent and progressing micro-
be different in BRAF-positive tumors in comparison to in other PTCs PTC? BRAF estimation may help to find the progression-potent
(Durante et al., 2007). tumors, but obviously at the present status of knowledge is not
To interpret these data, some more general remarks seem neces- powerful enough to predict the need for more aggressive treat-
sary: the down-expression of thyroid-specific genes is not specific ment than is presently recommended.
to BRAF-induced tumors, but is seen on the RNA level, the pro- 2. How to specify the more precise indications for lymphadenec-
tein level or both in every case of thyroid cancer (Ambroziak et tomy in DTC?
al., 2005; Huang et al., 2001; Krause et al., 2007; Lazar et al., 1999; Until now, the decision regarding lymphadenectomy is based
Skubis-Zegadlo et al., 2005). Especially, the direct attribution of the on the following rules: routine central lymphadenectomy and
poorer outcome of BRAF-positive PTC to the down-regulation of selective lateral lymphadenectomy meaning that the latter pro-
thyroid differentiation-related genes seems precocious. These two cedure is performed only when evidence or at least strong
facts are only associated, the causative relation is not proven. It suspicion exists for the presence of lymph node metastases
appears to be more aggressive tumor biology rather than poorer (Cooper et al., 2006; Pacini et al., 2006). Knowing the BRAF status
response to radioiodine therapy which causes poorer outcome in the primary tumor may prompt omission of central lym-
in BRAF-positive tumors. The impact of the lower expression of phadenectomy in BRAF-negative cases, however, this may lead
NIS protein in BRAF-mutated tumors is diminished by the fact to undertreatment of RET/PTC-positive young patients. On the
that both BRAF-positive and BRAF-negative tumors exhibit mainly other side, detection of a BRAF mutation would prompt a deci-
cytoplasmic localization of NIS, which precludes that protein’s sion to perform central lymphadenectomy in small tumors, a
functionality. The down-expression of NIS in BRAF-positive PTCs strategy that is not obligatory today, and hence could lead to
is seen also in microcarcinomas (Oler and Cerutti, 2009) and overtreatment.
even in these low-risk tumors, BRAF mutation has been related 3. How to optimize the decision for adjuvant radioiodine after
to development of more advanced disease, while the prognosis is surgery, especially in the “gray zone” of T1-T2N0-N1a patients?
excellent. Basing the decision on BRAF status would mean that about in
2/3 of the patients in this stage adjuvant treatment after surgery
5.1.3. BRAF mutation is associated with the more aggressive would be necessary. However, the indications for a standard
tumor phenotype in PTC adjuvant treatment with radioiodine are doubtful in this group
The exact molecular mechanisms of the invasive potential of in view of decreased expression of genes responsible for iodine
BRAF-induced tumors are still not known. As mentioned above, metabolism in BRAF-positive cancers.
the poorer outcome in patients bearing BRAF-positive tumors is 4. How to treat patients with advanced DTC?
with certainty not merely the result of poorer efficacy of radioio- We do not feel that the information on BRAF status changes
dine treatment, due to impaired radioiodine uptake and retention. anyhow the present treatment rules—even the success of
The prominent activation of the MEK–ERK1/2 pathway in BRAF- sorafenib, a tyrosine kinase inhibitor which is able to target BRAF
mutated thyrocytes is well known, but the cascade of subsequent as well, has not been shown dependent on BRAF mutation status
events is not well defined. Interestingly, MET overexpression, an (Kloos et al., 2009). On the other hand, data from other cancers
event related to poorer PTC prognosis, is seen mainly in BRAF- suggest that tumors bearing a BRAF mutation may be particularly
positive tumors (Giordano et al., 2005). Also, the mean Ki-67 index, prone to respond to MEK inhibitors (Ball et al., 2007), an obser-
an indicator of tumor cell proliferation, has been reported to be sig- vation that deserves confirmation in the context of advanced
nificantly higher in BRAF-positive than in BRAF-negative patients thyroid cancer.
D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 23

5.3. Conclusions from the transcriptome-wide analyses of DTCs titf1/nkx2-1 and pax-8 transcription factor genes in papillary thyroid cancer.
Thyroid 15 (10), 1137–1146.
Ball, D.W., Jin, N., Rosen, D.M., Dackiw, A., Sidransky, D., Xing, M., Nelkin, B.D., 2007.
1. Among the genetic factors analyzed for their prognostic signif- Selective growth inhibition in BRAF mutant thyroid cancer by the mitogen-
icance, the presence of activating BRAF mutation in papillary activated protein kinase kinase 1/2 inhibitor AZD6244. J. Clin. Endocrinol. Metab.
thyroid cancer seems the only one to be considered in clinical 92 (12), 4712–4718.
Balmanno, K., Cook, S.J., 2009. Tumour cell survival signalling by the ERK1/2 path-
practice. However, the putative significance of BRAF-positivity way. Cell Death Differ. 16 (3), 368–377.
is limited by the fact of its frequent occurrence. Further studies Bardet, S., Malville, E., Rame, J.P., Babin, E., Samama, G., De Raucourt, D., Michels, J.J.,
are necessary to stratifiy BRAF-positive papillary thyroid cancers Reznik, Y., Henry-Amar, M., 2008. Macroscopic lymph-node involvement and
neck dissection predict lymph-node recurrence in papillary thyroid carcinoma.
and to indentifiy molecular factors related to the subset with Eur. J. Endocrinol. 158 (4), 551–560.
poor prognosis. Baris, O., Mirebeau-Prunier, D., Savagner, F., Rodien, P., Ballester, B., Loriod, B.,
2. By gene signature, the difference between BRAF-positive and Granjeaud, S., Guyetant, S., Franc, B., Houlgatte, R., Reynier, P., Malthiery, Y., 2005.
Gene profiling reveals specific oncogenic mechanisms and signaling pathways
BRAF-negative PTC is so distinct that BRAF-positive cancer should in oncocytic and papillary thyroid carcinoma. Oncogene 24 (25), 4155–4161.
be regarded as a molecular subtype of PTC. There is an urgent Beasley, N.J., Lee, J., Eski, S., Walfish, P., Witterick, I., Freeman, J.L., 2002. Impact
need to investigate, which gene signature within the BRAF- of nodal metastases on prognosis in patients with well-differentiated thyroid
cancer. Arch. Otolaryngol. Head Neck Surg. 128 (7), 825–828.
related gene expression profile is specifically correlated with Bohm, J., Niskanen, L., Kiraly, K., Kellokoski, J., Eskelinen, M., Hollmen, S., Alhava,
poor prognosis. E., Kosma, V.M., 2000. Expression and prognostic value of alpha-, beta-, and
3. Differences between papillary and follicular cancers shown until gamma-catenins indifferentiated thyroid carcinoma. J. Clin. Endocrinol. Metab.
85 (12), 4806–4811.
now by gene expression profiling are profound, but they are
Bonnet, S, Hartl, D., Leboulleux, S., Baudin, E., Lumbroso, J.D., Al, G.A., Chami, L.,
still not sufficiently well characterized, thus, their relation to Schlumberger, M., Travagli, J.P., 2009. Prophylactic lymph node dissection for
the putative differences in prognosis between these histotypes papillary thyroid cancer less than 2 cm: implications for radioiodine treatment.
is precocious. J. Clin. Endocrinol. Metab. 94 (4), 1162–1167.
Brecelj, E., Frkovic, G.S., Auersperg, M., Bracko, M., 2005. Prognostic value of E-
cadherin expression in thyroid follicular carcinoma. Eur. J. Surg. Oncol. 31 (5),
Acknowledgments 544–548.
Brierley, J.D., Panzarella, T., Tsang, R.W., Gospodarowicz, M.K., O’Sullivan, B., 1997.
A comparison of different staging systems predictability of patient outcome.
The authors thank Robert J. Marlowe for very engaged and com- Thyroid carcinoma as an example. Cancer 79 (12), 2414–2423.
petent editorial assistance and Emilia Wilk and Monika Kowal for Brzezianska, E., Pastuszak-Lewandoska, D., Lewinski, A., 2007. Rearrangements of
NTRK1 oncogene in papillary thyroid carcinoma. Neuroendocrinol. Lett. 28 (3),
excellent general assistance. Daria Handkiewicz-Junak was partly
221–229.
supported by Polish Ministry of Science and Higher Education Cady, B., Rossi, R., 1988. An expanded view of risk-group definition in differentiated
grant no. N40110132/2138 and Barbara Jarzab was supported by thyroid carcinoma. Surgery 104 (6), 947–953.
European FP6 project no. 036495, acronym: Genrisk-T and Polish Canzian, F., Amati, P., Harach, H.R., Kraimps, J.L., Lesueur, F., Barbier, J., Levillain, P.,
Romeo, G., Bonneau, D., 1998. A gene predisposing to familial thyroid tumors
Ministry of Science and Higher Education grant no. PBZ-MNiI- with cell oxyphilia maps to chromosome 19p13.2. Am. J. Hum. Genet. 63 (6),
2/1/2005. 1743–1748.
Carver, B.S., Tran, J., Gopalan, A., Chen, Z., Shaikh, S., Carracedo, A., Alimonti, A.,
Nardella, C., Varmeh, S., Scardino, P.T., Cordon-Cardo, C., Gerald, W., Pandolfi,
References P.P., 2009. Aberrant ERG expression cooperates with loss of PTEN to promote
cancer progression in the prostate. Nat. Genet. 41 (5), 619–624.
Abrosimov, A., Saenko, V., Meirmanov, S., Nakashima, M., Rogounovitch, T., Shkurko, Castro, P., Rebocho, A.P., Soares, R.J., Magalhaes, J., Roque, L., Trovisco, V., Vieira,
O., Lushnikov, E., Mitsutake, N., Namba, H., Yamashita, S., 2007a. The cytoplas- d.C.I., Cardoso-de-Oliveira, M., Fonseca, E., Soares, P., Sobrinho-Simoes, M., 2006.
mic expression of MUC1 in papillary thyroid carcinoma of different histological PAX8–PPARgamma rearrangement is frequently detected in the follicular vari-
variants and its correlation with cyclin D1 overexpression. Endocr. Pathol. 18 ant of papillary thyroid carcinoma. J. Clin. Endocrinol. Metab. 91 (1), 213–220.
(2), 68–75. Cavaco, B.M., Batista, P.F., Martins, C., Banito, A., do Rosário,.F., Limbert, E., Sobrinho,
Abrosimov, A., Saenko, V., Rogounovitch, T., Namba, H., Lushnikov, E., Mitsutake, N., L.G., Leite, V., 2008. Familial non-medullary thyroid carcinoma (FNMTC): analy-
Yamashita, S., 2007b. Different structural components of conventional papillary sis of fPTC/PRN, NMTC1, MNG1 and TCO susceptibility loci and identification of
thyroid carcinoma display mostly identical BRAF status. Int. J. Cancer 120 (1), somatic BRAF and RAS mutations. Endocr. Relat. Cancer 15 (1), 207–215.
196–200. Cheung, L., Messina, M., Gill, A., Clarkson, A., Learoyd, D., Delbridge, L., Wentworth,
Abubaker, J., Jehan, Z., Bavi, P., Sultana, M., Al-Harbi, S., Ibrahim, M., Al-Nuaim, A., J., Philips, J., Clifton-Bligh, R., Robinson, B.G., 2003. Detection of the PAX8–PPAR
Ahmed, M., Amin, T., Al-Fehaily, M., Al-Sanea, O., Al-Dayel, F., Uddin, S., Al- gamma fusion oncogene in both follicular thyroid carcinomas and adenomas. J.
Kuraya, K.S., 2008. Clinicopathological analysis of papillary thyroid cancer with Clin. Endocrinol. Metab. 88 (1), 354–357.
PIK3CA alterations in a Middle Eastern population. J. Clin. Endocrinol. Metab. 93 Chevillard, S., Ugolin, N., Vielh, P., Ory, K., Levalois, C., Elliott, D., Clayman,
(2), 611–618. G.L., El-Naggar, A.K., 2004. Gene expression profiling of differentiated thy-
Adeniran, A.J., Zhu, Z., Gandhi, M., Steward, D.L., Fidler, J.P., Giordano, T.J., Biddinger, roid neoplasms: diagnostic and clinical implications. Clin. Cancer Res. 10 (19),
P.W., Nikiforov, Y.E., 2006. Correlation between genetic alterations and micro- 6586–6597.
scopic features, clinical manifestations, and prognostic characteristics of thyroid Clark, O.H., Duh, Q.Y., 1991. Thyroid cancer. Med. Clin. North Am. 75 (1), 211–234.
papillary carcinomas. Am. J. Surg. Pathol. 30 (2), 216–222. Collins, B.J., Chiappetta, G., Schneider, A.B., Santoro, M., Pentimalli, F., Fogelfeld, L.,
Aherne, S.T., Smyth, P.C., Flavin, R.J., Russell, S.M., Denning, K.M., Li, J.H., Guenther, Gierlowski, T., Shore-Freedman, E., Jaffe, G., Fusco, A., 2002. RET expression in
S.M., O’Leary, J.J., Sheils, O.M., 2008. Geographical mapping of a multifocal thy- papillary thyroid cancer from patients irradiated in childhood for benign condi-
roid tumour using genetic alteration analysis & miRNA profiling. Mol. Cancer 7, tions. J. Clin. Endocrinol. Metab. 87 (8), 3941–3946.
89. Cooper, D.S., Doherty, G.M., Haugen, B.R., Kloos, R.T., Lee, S.L., Mandel, S.J., Mazza-
Akslen, L.A., 1993. Prognostic importance of histologic grading in papillary thyroid ferri, E.L., McIver, B., Sherman, S.I., Tuttle, R.M., 2006. Management guidelines
carcinoma. Cancer 72 (9), 2680–2685. for patients with thyroid nodules and differentiated thyroid cancer. Thyroid 16
Akslen, L.A., Varhaug, J.E., 1995. Oncoproteins and tumor progression in papillary (2), 109–142.
thyroid carcinoma: presence of epidermal growth factor receptor, c-erbB-2 Cooper, D.S., Doherty, G.M., Haugen, B.R., Kloos, R.T., Lee, S.L., Mandel, S.J., Mazzaferri,
protein, estrogen receptor related protein, p21-ras protein, and proliferation E.L., McIver, B., Pacini, F., Schlumberger, M., Sherman, S.I., Steward, D.L., Tuttle,
indicators in relation to tumor recurrences and patient survival. Cancer 76 (9), R.M., 2009. Revised American Thyroid Association management guidelinesfor
1643–1654. patients with thyroid nodules and differentiated thyroid cancer. Thyroid 19 (11),
Aldred, M.A, Ginn-Pease, M.E., Morrison, C.D., Popkie, A.P., Gimm, O., Hoang-Vu, 1167–1214.
C., Krause, U., Dralle, H., Jhiang, S.M., Plass, C., Eng, C., 2003. Caveolin-1 and Costa, A.M., Herrero, A., Fresno, M.F., Heymann, J., Alvarez, J.A., Cameselle-Teijeiro,
caveolin-2, together with three bone morphogenetic protein-related genes, may J., Garcia-Rostan, G., 2008. BRAF mutation associated with other genetic events
encode novel tumor suppressors down-regulated in sporadic follicular thyroid identifies a subset of aggressive papillary thyroid carcinoma. Clin. Endocrinol.
carcinogenesis. Cancer Res. 63 (11), 2864–2871. (Oxf.) 68 (4), 618–634.
Aldred, M.A., Huang, Y., Liyanarachchi, S., Pellegata, N.S., Gimm, O., Jhiang, S., Czarniecka, A., 2004. Indication for Less Than Total Thyroidectomies in Differentiated
Davuluri, R.V., de la Chapelle, A., Eng, C., 2004. Papillary and follicular thyroid Thyroid Cancer. Medical Unierrsity of Silesia, Zabrze, Poland (Ref Type: Generic).
carcinomas show distinctly different microarray expression profiles and can be D’Avanzo, A., Ituarte, P., Treseler, P., Kebebew, E., Wu, J., Wong, M., Duh, Q.Y.,
distinguished by a minimum of five genes. J. Clin. Oncol. 22 (17), 3531–3539. Siperstein, A.E., Clark, O.H., 2004. Prognostic scoring systems in patients with
Ambroziak, M., Pachucki, J., Stachlewska-Nasfeter, E., Nauman, J., Nauman, A., 2005. follicular thyroid cancer: a comparison of different staging systems in predicting
Disturbed expression of type 1 and type 2 iodothyronine deiodinase as well as the patient outcome. Thyroid 14 (6), 453–458.
24 D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28

Dankort, D., Filenova, E., Collado, M., Serrano, M., Jones, K., McMahon, M., 2007. A new Frisk, T., Foukakis, T., Dwight, T., Lundberg, J., Hoog, A., Wallin, G., Eng, C., Zedenius,
mouse model to explore the initiation, progression, and therapy of BRAFV600E- J., Larsson, C., 2002. Silencing of the PTEN tumor-suppressor gene in anaplastic
induced lung tumors. Genes Dev. 21 (4), 379–384. thyroid cancer. Genes Chromosomes Cancer 35 (1), 74–80.
Davies, H, Bignell, G.R., Cox, C., Stephens, P., Edkins, S., Clegg, S., Teague, J., Woffendin, Fugazzola, L., Mannavola, D., Cirello, V., Vannucchi, G., Muzza, M., Vicentini, L., Beck-
H., Garnett, M.J., Bottomley, W., Davis, N., Dicks, E., Ewing, R., Floyd, Y., Gray, K., Peccoz, P., 2004. BRAF mutations in an Italian cohort of thyroid cancers. Clin.
Hall, S., Hawes, R., Hughes, J., Kosmidou, V., Menzies, A., Mould, C., Parker, A., Endocrinol. (Oxf.) 61 (2), 239–243.
Stevens, C., Watt, S., Hooper, S., Wilson, R., Jayatilake, H., Gusterson, B.A., Cooper, Gandhi, M., Medvedovic, M., Stringer, J.R., Nikiforov, Y.E., 2006. Interphase chromo-
C., Shipley, J., Hargrave, D., Pritchard-Jones, K., Maitland, N., Chenevix-Trench, G., some folding determines spatial proximity of genes participating in carcinogenic
Riggins, G.J., Bigner, D.D., Palmieri, G., Cossu, A., Flanagan, A., Nicholson, A., Ho, RET/PTC rearrangements. Oncogene 25 (16), 2360–2366.
J.W., Leung, S.Y., Yuen, S.T., Weber, B.L., Seigler, H.F., Darrow, T.L., Paterson, H., Garcia-Rostan, G., Camp, R.L., Herrero, A., Carcangiu, M.L., Rimm, D.L., Tallini,
Marais, R., Marshall, C.J., Wooster, R., Stratton, M.R., Futreal, P.A., 2002. Mutations G., 2001. Beta-catenin dysregulation in thyroid neoplasms: down-regulation,
of the BRAF gene in human cancer. Nature 417 (6892), 949–954. aberrant nuclear expression, and CTNNB1 exon 3 mutations are markers for
DeGroot, L.J., Kaplan, E.L., McCormick, M., Straus, F.H., 1990. Natural history, treat- aggressive tumor phenotypes and poor prognosis. Am. J. Pathol. 158 (3),
ment, and course of papillary thyroid carcinoma. J. Clin. Endocrinol. Metab. 71 987–996.
(2), 414–424. Garcia-Rostan, G., Costa, A.M., Pereira-Castro, I., Salvatore, G., Hernandez, R.,
DeLuca, A.M., Srinivas, A., Alani, R.M., 2008. BRAF kinase in melanoma development Hermsem, M.J., Herrero, A., Fusco, A., Cameselle-Teijeiro, J., Santoro, M., 2005.
and progression. Expert Rev. Mol. Med. 10, e6. Mutation of the PIK3CA gene in anaplastic thyroid cancer. Cancer Res. 65 (22),
Di Renzo, M.F., Olivero, M., Ferro, S., Prat, M., Bongarzone, I., Pilotti, S., Belfiore, A., 10199–10207.
Costantino, A., Vigneri, R., Pierotti, M.A., 1992. Overexpression of the c-MET/HGF Garcia-Rostan, G., Tallini, G., Herrero, A., D’Aquila, T.G., Carcangiu, M.L., Rimm, D.L.,
receptor gene in human thyroid carcinomas. Oncogene 7 (12), 2549–2553. 1999. Frequent mutation and nuclear localization of beta-catenin in anaplastic
Durante, C., Haddy, N., Baudin, E., Leboulleux, S., Hartl, D., Travagli, J.P., Caillou, B., thyroid carcinoma. Cancer Res. 59 (8), 1811–1815.
Ricard, M., Lumbroso, J.D., De Vathaire, F., Schlumberger, M., 2006. Long-term Garcia-Rostan, G., Zhao, H., Camp, R.L., Pollan, M., Herrero, A., Pardo, J., Wu, R.,
outcome of 444 patients with distant metastases from papillary and follicular Carcangiu, M.L., Costa, J., Tallini, G., 2003. ras mutations are associated with
thyroid carcinoma: benefits and limits of radioiodine therapy. J. Clin. Endocrinol. aggressive tumor phenotypes and poor prognosis in thyroid cancer. J. Clin. Oncol.
Metab. 91 (8), 2892–2899. 21 (17), 3226–3235.
Durante, C., Puxeddu, E., Ferretti, E., Morisi, R., Moretti, S., Bruno, R., Barbi, F., Avenia, Garnett, M.J., Marais, R., 2004. Guilty as charged: B-RAF is a human oncogene. Cancer
N., Scipioni, A., Verrienti, A., Tosi, E., Cavaliere, A., Gulino, A., Filetti, S., Russo, D., Cell 6 (4), 313–319.
2007. BRAF mutations in papillary thyroid carcinomas inhibit genes involved in Ghossein, R., LiVolsi, V.A., 2008. Papillary thyroid carcinoma tall cell variant. Thyroid
iodine metabolism. J. Clin. Endocrinol. Metab. 92 (7), 2840–2843. 18 (11), 1179–1181.
Elisei, R., Ugolini, C., Viola, D., Lupi, C., Biagini, A., Giannini, R., Romei, C., Miccoli, P., Gimm, O., Perren, A., Weng, L.P., Marsh, D.J., Yeh, J.J., Ziebold, U., Gil, E., Hinze, R.,
Pinchera, A., Basolo, F., 2008. BRAF(V600E) mutation and outcome of patients Delbridge, L., Lees, J.A., Mutter, G.L., Robinson, B.G., Komminoth, P., Dralle, H.,
with papillary thyroid carcinoma: a 15-year median follow-up study. J. Clin. Eng, C., 2000. Differential nuclear and cytoplasmic expression of PTEN in normal
Endocrinol. Metab. 93 (10), 3943–3949. thyroid tissue, and benign and malignant epithelial thyroid tumors. Am. J. Pathol.
Esapa, C.T., Johnson, S.J., Kendall-Taylor, P., Lennard, T.W., Harris, P.E., 1999. Preva- 156 (5), 1693–1700.
lence of Ras mutations in thyroid neoplasia. Clin. Endocrinol. (Oxf.) 50 (4), Giordano, T.J., 2008. Genome-wide studies in thyroid neoplasia. Endocrinol. Metab.
529–535. Clin. North Am. 37 (2), 311-viii.
Eszlinger, M., Krohn, K., Hauptmann, S., Dralle, H., Giordano, T.J., Paschke, R., 2008. Giordano, T.J., Au, A.Y., Kuick, R., Thomas, D.G., Rhodes, D.R., Wilhelm Jr., K.G.,
Perspectives for improved and more accurate classification of thyroid epithelial Vinco, M., Misek, D.E., Sanders, D., Zhu, Z., Ciampi, R., Hanash, S., Chinnaiyan, A.,
tumors. J. Clin. Endocrinol. Metab. 93 (9), 3286–3294. Clifton-Bligh, R.J., Robinson, B.G., Nikiforov, Y.E., Koenig, R.J., 2006. Delineation,
Eustatia-Rutten, C.F., Corssmit, E.P., Biermasz, N.R., Pereira, A.M., Romijn, J.A., Smit, functional validation, and bioinformatic evaluation of gene expression in thy-
J.W., 2006. Survival and death causes in differentiated thyroid carcinoma. J. Clin. roid follicular carcinomas with the PAX8–PPARG translocation. Clin. Cancer Res.
Endocrinol. Metab. 91 (1), 313–319. 12 (7 Pt 1), 1983–1993.
Fagin, J.A., 2004. Challenging dogma in thyroid cancer molecular genetics—role Giordano, T.J., Kuick, R., Thomas, D.G., Misek, D.E., Vinco, M., Sanders, D., Zhu, Z.,
of RET/PTC and BRAF in tumor initiation. J. Clin. Endocrinol. Metab. 89 (9), Ciampi, R., Roh, M., Shedden, K., Gauger, P., Doherty, G., Thompson, N.W., Hanash,
4264–4266. S., Koenig, R.J., Nikiforov, Y.E., 2005. Molecular classification of papillary thyroid
Fagin, J.A., Mitsiades, N., 2008. Molecular pathology of thyroid cancer: diagnostic carcinoma: distinct BRAF, RAS, and RET/PTC mutation-specific gene expression
and clinical implications. Best Pract. Res. Clin. Endocrinol. Metab. 22 (6), 955– profiles discovered by DNA microarray analysis. Oncogene 24 (44), 6646–6656.
969. Goutas, N., Vlachodimitropoulos, D., Bouka, M., Lazaris, A.C., Nasioulas, G., Gazouli,
Fenton, C.L., Lukes, Y., Nicholson, D., Dinauer, C.A., Francis, G.L., Tuttle, R.M., 2000. M., 2008. BRAF and K-RAS mutation in a Greek papillary and medullary thyroid
The ret/PTC mutations are common in sporadic papillary thyroid carcinoma of carcinoma cohort. Anticancer Res. 28 (1A), 305–308.
children and young adults. J. Clin. Endocrinol. Metab. 85 (3), 1170–1175. Greene, F.L., Page, D.L., Fleming, I.D., Fritz, A., 2002. AJCC Cancer Staging Manual, 6th
Ferby, I., Reschke, M., Kudlacek, O., Knyazev, P., Pante, G., Amann, K., Sommergruber, edition. Springer.
W., Kraut, N., Ullrich, A., Fassler, R., Klein, R., 2006. Mig6 is a negative regulator Gregory, P.J., Wang, X., Allard, B.L., Sahin, M., Wang, X.L., Hay, I.D., Hiddinga,
of EGF receptor-mediated skin morphogenesis and tumor formation. Nat. Med. H.J., Deshpande, S.S., Kroll, T.G., Grebe, S.K., Eberhardt, N.L., McIver, B., 2004.
12 (5), 568–573. The PAX8/PPARgamma fusion oncoprotein transforms immortalized human
Ferenc, T., Lewinski, A., Lange, D., Niewiadomska, H., Sygut, J., Sporny, S., Jarzab, B., thyrocytes through a mechanism probably involving wild-type PPARgamma
Satacinska-Los, E., Kulig, A., Wloch, J., 2005. Analysis of cyclin D1 and retinoblas- inhibition. Oncogene 23 (20), 3634–3641.
toma protein immunoreactivity in follicular thyroid tumors. Pol. J. Pathol. 56 (1), Griffith, O.L., Melck, A., Jones, S.J., Wiseman, S.M., 2006. Meta-analysis and meta-
27–35. review of thyroid cancer gene expression profiling studies identifies important
Figge, J., del Rosario, A.D., Gerasimov, G., Dedov, I., Bronstein, M., Troshina, K., Alexan- diagnostic biomarkers. J Clin Oncol. 24 (31), 5043–5051.
drova, G., Kallakury, B.V., Bui, H.X., Bratslavsky, G., 1994. Preferential expression Guan, H., Ji, M., Bao, R., Yu, H., Wang, Y., Hou, P., Zhang, Y., Shan, Z., Teng, W., Xing,
of the cell adhesion molecule CD44 in papillary thyroid carcinoma. Exp. Mol. M., 2009. Association of high iodine intake with the T1799A BRAF mutation in
Pathol. 61 (3), 203–211. papillary thyroid cancer. J. Clin. Endocrinol. Metab. 94 (5), 1612–1617.
Finley, D.J., Arora, N., Zhu, B., Gallagher, L., Fahey III, T.J., 2004a. Molecular pro- Guarino, V., Faviana, P., Salvatore, G., Castellone, M.D., Cirafici, A.M., De Falco, V.,
filing distinguishes papillary carcinoma from benign thyroid nodules. J. Clin. Celetti, A., Giannini, R., Basolo, F., Melillo, R.M., Santoro, M., 2005. Osteopontin
Endocrinol. Metab. 89 (7), 3214–3223. is overexpressed in human papillary thyroid carcinomas and enhances thyroid
Finley, D.J., Zhu, B., Barden, C.B., Fahey III, T.J., 2004b. Discrimination of benign carcinoma cell invasiveness. J. Clin. Endocrinol. Metab. 90 (9), 5270–5278.
and malignant thyroid nodules by molecular profiling. Ann. Surg. 240 (3), 425– Gudmundsson, J., Sulem, P., Gudbjartsson, D.F., Jonasson, J.G., Sigurdsson, A., Bergth-
436. orsson, J.T., He, H., Blondal, T., Geller, F., Jakobsdottir, M., Magnusdottir, D.N.,
Franc, B., de la, S.P., Lange, F., Hoang, C., Louvel, A., de Roquancourt, A., Vilde, F., Matthiasdottir, S., Stacey, S.N., Skarphedinsson, O.B., Helgadottir, H., Li, W., Nagy,
Hejblum, G., Chevret, S., Chastang, C., 2003. Interobserver and intraobserver R., Aguillo, E., Faure, E., Prats, E., Saez, B., Martinez, M., Eyjolfsson, G.I., Bjorns-
reproducibility in the histopathology of follicular thyroid carcinoma. Hum. dottir, U.S., Holm, H., Kristjansson, K., Frigge, M.L., Kristvinsson, H., Gulcher,
Pathol. 34 (11), 1092–1100. J.R., Jonsson, T., Rafnar, T., Hjartarsson, H., Mayordomo, J.I., de la Chapelle, A.,
Franzoni, A., Dima, M., D’Agostino, M., Puppin, C., Fabbro, D., Loreto, C.D., Pandolfi, Hrafnkelsson, J., Thorsteinsdottir, U., Kong, A., Stefansson, K., 2009. Common
M., Puxeddu, E., Moretti, S., Celano, M., Bruno, R., Filetti, S., Russo, D., Damante, variants on 9q22.33 and 14q13.3 predispose to thyroid cancer in European pop-
G., 2009. Prohibitin is overexpressed in papillary thyroid carcinomas bearing the ulations. Nat. Genet. 41 (4), 460–464.
BRAF(V600E) mutation. Thyroid 19 (3), 247–255. Gulcelik, M.A., Gulcelik, N.E., Kuru, B., Camlibel, M., Alagol, H., 2007. Prognostic fac-
Frasca, F., Nucera, C., Pellegriti, G., Gangemi, P., Attard, M., Stella, M., Loda, M., tors determining survival in differentiated thyroid cancer. J. Surg. Oncol. 96 (7),
Vella, V., Giordano, C., Trimarchi, F., Mazzon, E., Belfiore, A., Vigneri, R., 2008. 598–604.
BRAF(V600E) mutation and the biology of papillary thyroid cancer. Endocr. Relat. Haferlach, T., Kohlmann, A., Kern, W., Hiddemann, W., Schnittger, S., Schoch, C., 2003.
Cancer 15 (1), 191–205. Gene expression profiling as a tool for the diagnosis of acute leukemias. Semin.
Frattini, M., Ferrario, C., Bressan, P., Balestra, D., De Cecco, L., Mondellini, P., Bon- Hematol. 40 (4), 281–295.
garzone, I., Collini, P., Gariboldi, M., Pilotti, S., Pierotti, M.A., Greco, A., 2004. Hamada, A., Mankovskaya, S., Saenko, V., Rogounovitch, T., Mine, M., Namba, H.,
Alternative mutations of BRAF, RET and NTRK1 are associated with similar but Nakashima, M., Demidchik, Y., Demidchik, E., Yamashita, S., 2005. Diagnostic
distinct gene expression patterns in papillary thyroid cancer. Oncogene 23 (44), usefulness of PCR profiling of the differentially expressed marker genes in thy-
7436–7440. roid papillary carcinomas. Cancer Lett. 224 (2), 289–301.
D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 25

Handkiewcz-Junak, D., Banasik, T., Kolosza, Z., Roskosz, J., Kukulska, A., Puch, Z., H.R., Sherman, S.I., 2006. Outcomes of patients with differentiated thyroid car-
Jarzab, B., 2006. Risk of malignant tumors in first-degree relatives of patients cinoma following initial therapy. Thyroid 16 (12), 1229–1242.
with differentiated thyroid cancer—a hospital based study. Neoplasma 53 (1), Jubb, A.M., Pham, T.Q., Hanby, A.M., Frantz, G.D., Peale, F.V., Wu, T.D., Koeppen,
67–72. H.W., Hillan, K.J., 2004. Expression of vascular endothelial growth factor, hypoxia
Handkiewicz-Junak, D., Wloch, J., Roskosz, J., Krajewska, J., Kropinska, A., Pomorski, inducible factor 1alpha, and carbonic anhydrase IX in human tumours. J. Clin.
L., Kukulska, A., Prokurat, A., Wygoda, Z., Jarzab, B., 2007. Total thyroidectomy Pathol. 57 (5), 504–512.
and adjuvant radioiodine treatment independently decrease locoregional recur- Kakudo, K., Tang, W., Ito, Y., Mori, I., Nakamura, Y., Miyauchi, A., 2004. Papillary
rence risk in childhood and adolescent differentiated thyroid cancer. J. Nucl. carcinoma of the thyroid in Japan: subclassification of common type and iden-
Med. 48 (6), 879–888. tification of low risk group. J. Clin. Pathol. 57 (10), 1041–1046.
Harwood, J., Clark, O.H., Dunphy, J.E., 1978. Significance of lymph node metastasis Kapran, Y., Ozbey, N., Molvalilar, S., Sencer, E., Dizdaroglu, F., Ozarmagan, S., 2002.
in differentiated thyroid cancer. Am. J. Surg. 136 (1), 107–112. Immunohistochemical detection of E-cadherin, alpha- and beta-catenins in pap-
Hay, I.D., Bergstralh, E.J., Goellner, J.R., Ebersold, J.R., Grant, C.S., 1993. Predicting illary thyroid carcinoma. J. Endocrinol. Invest. 25 (7), 578–585.
outcome in papillary thyroid carcinoma: development of a reliable prognostic Kashima, K., Yokoyama, S., Noguchi, S., Murakami, N., Yamashita, H., Watanabe, S.,
scoring system in a cohort of 1779 patients surgically treated at one institution Uchino, S., Toda, M., Sasaki, A., Daa, T., Nakayama, I., 1998. Chronic thyroiditis
during 1940 through 1989. Surgery 114 (6), 1050–1057. as a favorable prognostic factor in papillary thyroid carcinoma. Thyroid 8 (3),
Hay, I.D., Grant, C.S., Taylor, W.F., McConahey, W.M., 1987. Ipsilateral lobectomy 197–202.
versus bilateral lobar resection in papillary thyroid carcinoma: a retrospective Kebebew, E., Weng, J., Bauer, J., Ranvier, G., Clark, O.H., Duh, Q.Y., Shibru, D., Bastian,
analysis of surgical outcome using a novel prognostic scoring system. Surgery B., Griffin, A., 2007. The prevalence and prognostic value of BRAF mutation in
102 (6), 1088–1095. thyroid cancer. Ann. Surg. 246 (3), 466–470.
Henderson, Y.C., Shellenberger, T.D., Williams, M.D., El-Naggar, A.K., Fredrick, M.J., Khoo, M.L., Beasley, N.J., Ezzat, S., Freeman, J.L., Asa, S.L., 2002. Overexpression of
Cieply, K.M., Clayman, G.L., 2009. High rate of BRAF and RET/PTC dual mutations cyclin D1 and underexpression of p27 predict lymph node metastases in papil-
associated with recurrent papillary thyroid carcinoma. Clin. Cancer Res. 15 (2), lary thyroid carcinoma. J. Clin. Endocrinol. Metab. 87 (4), 1814–1818.
485–491. Kim, J., Giuliano, A.E., Turner, R.R., Gaffney, R.E., Umetani, N., Kitago, M., Elashoff,
Hilger, R.A., Scheulen, M.E., Strumberg, D., 2002. The Ras–Raf–MEK–ERK pathway in D., Hoon, D.S., 2006a. Lymphatic mapping establishes the role of BRAF gene
the treatment of cancer. Onkologie 25 (6), 511–518. mutation in papillary thyroid carcinoma. Ann. Surg. 244 (5), 799–804.
Hoos, A., Stojadinovic, A., Singh, B., Dudas, M.E., Leung, D.H., Shaha, A.R., Shah, Kim, K.H., Suh, K.S., Kang, D.W., Kang, D.Y., 2005. Mutations of the BRAF gene in
J.P., Brennan, M.F., Cordon-Cardo, C., Ghossein, R., 2002. Clinical significance papillary thyroid carcinoma and in Hashimoto’s thyroiditis. Pathol. Int. 55 (9),
of molecular expression profiles of Hurthle cell tumors of the thyroid gland 540–545.
analyzed via tissue microarrays. Am. J. Pathol. 160 (1), 175–183. Kim, S.K., Song, K.H., Lim, S.D., Lim, Y.C., Yoo, Y.B., Kim, J.S., Hwang, T.S., 2009. Clinical
Hou, P., Liu, D., Shan, Y., Hu, S., Studeman, K., Condouris, S., Wang, Y., Trink, A., El- and pathological features and the BRAF(V600E) mutation in patients with pap-
Naggar, A.K., Tallini, G., Vasko, V., Xing, M., 2007. Genetic alterations and their illary thyroid carcinoma with and without concurrent Hashimoto thyroiditis.
relationship in the phosphatidylinositol 3-kinase/Akt pathway in thyroid cancer. Thyroid 19 (2), 137–141.
Clin. Cancer Res. 13 (4), 1161–1170. Kim, T.Y., Kim, W.B., Rhee, Y.S., Song, J.Y., Kim, J.M., Gong, G., Lee, S., Kim, S.Y., Kim,
Hu, J., Zhang, X., Nothnick, W.B., Spencer, T.E., 2004. Matrix metalloproteinases and S.C., Hong, S.J., Shong, Y.K., 2006b. The BRAF mutation is useful for prediction
their tissue inhibitors in the developing neonatal mouse uterus. Biol. Reprod. 71 of clinical recurrence in low-risk patients with conventional papillary thyroid
(5), 1598–1604. carcinoma. Clin. Endocrinol. (Oxf.) 65 (3), 364–368.
Huang, Y., Prasad, M., Lemon, W.J., Hampel, H., Wright, F.A., Kornacker, K., LiVolsi, Kimura, E.T., Nikiforova, M.N., Zhu, Z., Knauf, J.A., Nikiforov, Y.E., Fagin, J.A., 2003.
V., Frankel, W., Kloos, R.T., Eng, C., Pellegata, N.S., de la Chapelle, A., 2001. Gene High prevalence of BRAF mutations in thyroid cancer: genetic evidence for con-
expression in papillary thyroid carcinoma reveals highly consistent profiles. stitutive activation of the RET/PTC–RAS–BRAF signaling pathway in papillary
Proc. Natl. Acad. Sci. U.S.A. 98 (26), 15044–15049. thyroid carcinoma. Cancer Res. 63 (7), 1454–1457.
Ito, Y., Hirokawa, M., Uruno, T., Kihara, M., Higashiyama, T., Takamura, Y., Miya, Klein, M., Vignaud, J.M., Hennequin, V., Toussaint, B., Bresler, L., Plenat, F., Leclere, J.,
A., Kobayashi, K., Matsuzuka, F., Miyauchi, A., 2008. Prevalence and biological Duprez, A., Weryha, G., 2001. Increased expression of the vascular endothelial
behaviour of variants of papillary thyroid carcinoma: experience at a single growth factor is a pejorative prognosis marker in papillary thyroid carcinoma.
institute. Pathology 40 (6), 617–622. J. Clin. Endocrinol. Metab. 86 (2), 656–658.
Ito, Y., Uruno, T., Takamura, Y., Miya, A., Kobayashi, K., Matsuzuka, F., Kuma, K., Kloos, R.T., Ringel, M.D., Knopp, M.V., Hall, N.C., King, M., Stevens, R., Liang, J., Wakely
Miyauchi, A., 2005. Papillary microcarcinomas of the thyroid with preopera- Jr., P.E., Vasko, V.V., Saji, M., Rittenberry, J., Wei, L., Arbogast, D., Collamore, M.,
tively detectable lymph node metastasis show significantly higher aggressive Wright, J.J., Grever, M., Shah, M.H., 2009. Phase II trial of sorafenib in metastatic
characteristics on immunohistochemical examination. Oncology 68 (2–3), thyroid cancer. J. Clin. Oncol. 27 (10), 1675–1684.
87–96. Knauf, J.A., Fagin, J.A., 2009. Role of MAPK pathway oncoproteins in thyroid cancer
Ito, Y., Yoshida, H., Maruo, R., Morita, S., Takano, T., Hirokawa, M., Yabuta, T., pathogenesis and as drug targets. Curr. Opin. Cell Biol. 21 (2), 296–303.
Fukushima, M., Inoue, H., Tomoda, C., Kihara, M., Uruno, T., Higashiyama, T., Knauf, J.A., Kuroda, H., Basu, S., Fagin, J.A., 2003a. RET/PTC-induced dedifferentiation
Takamura, Y., Miya, A., Kobayashi, K., Matsuzuka, F., Miyauchi, A., 2009. BRAF of thyroid cells is mediated through Y1062 signaling through SHC–RAS–MAP
mutation in papillary thyroid carcinoma in a Japanese population: its lack of kinase. Oncogene 22 (28), 4406–4412.
correlation with high-risk clinicopathological features and disease-free survival Knauf, J.A., Ma, X., Smith, E.P., Zhang, L., Mitsutake, N., Liao, X.H., Refetoff, S., Niki-
of patients. Endocr. J. 56 (1), 89–97. forov, Y.E., Fagin, J.A., 2005. Targeted expression of BRAFV600E in thyroid cells
Jabbour, E., Cortes, J.E., Kantarjian, H.M., 2008. Molecular monitoring in chronic of transgenic mice results in papillary thyroid cancers that undergo dedifferen-
myeloid leukemia: response to tyrosine kinase inhibitors and prognostic impli- tiation. Cancer Res. 65 (10), 4238–4245.
cations. Cancer 112 (10), 2112–2118. Knauf, J.A., Ouyang, B., Croyle, M., Kimura, E., Fagin, J.A., 2003b. Acute expression of
Jarzab, B., Handkiewicz-Junak, D., Wloch, J., 2005a. Juvenile differentiated thyroid RET/PTC induces isozyme-specific activation and subsequent downregulation
carcinoma and the role of radioiodine in its treatment: a qualitative review. of PKCepsilon in PCCL3 thyroid cells. Oncogene 22 (44), 6830–6838.
Endocr. Relat. Cancer 12 (4), 773–803. Kondo, T., Ezzat, S., Asa, S.L., 2006. Pathogenetic mechanisms in thyroid follicular-cell
Jarzab, B., Wiench, M., Fujarewicz, K., Simek, K., Jarzab, M., Oczko-Wojciechowska, neoplasia. Nat. Rev. Cancer 6 (4), 292–306.
M., Wloch, J., Czarniecka, A., Chmielik, E., Lange, D., Pawlaczek, A., Szpak, S., Krause, K., Karger, S., Gimm, O., Sheu, S.Y., Dralle, H., Tannapfel, A., Schmid, K.W.,
Gubala, E., Swierniak, A., 2005b. Gene expression profile of papillary thyroid Dupuy, C., Fuhrer, D., 2007. Characterisation of DEHAL1 expression in thyroid
cancer: sources of variability and diagnostic implications. Cancer Res. 65 (4), pathologies. Eur. J. Endocrinol. 156 (3), 295–301.
1587–1597. Kroll, T.G., Sarraf, P., Pecciarini, L., Chen, C.J., Mueller, E., Spiegelman, B.M., Fletcher,
Jazdzewski, K., Murray, E.L., Franssila, K., Jarzab, B., Schoenberg, D.R., de la Chapelle, J.A., 2000. PAX8–PPARgamma1 fusion oncogene in human thyroid carcinoma.
A., 2008. Common SNP in pre-miR-146a decreases mature miR expression and Science 289 (5483), 1357–1360.
predisposes to papillary thyroid carcinoma. Proc. Natl. Acad. Sci. U.S.A. 105 (20), Kremenevskaja, N., von, W.R., Rao, A.S., Schofl, C., Andersson, T., Brabant, G., 2005.
7269–7274. Wnt-5a has tumor suppressor activity in thyroid carcinoma. Oncogene 24 (13),
Jhiang, S.M., Sagartz, J.E., Tong, Q., Parker-Thornburg, J., Capen, C.C., Cho, J.Y., Xing, S., 2144–2154.
Ledent, C., 1996. Targeted expression of the ret/PTC1 oncogene induces papillary Kremser, R., Obrist, P., Spizzo, G., Erler, H., Kendler, D., Kemmler, G., Mikuz, G.,
thyroid carcinomas. Endocrinology 137 (1), 375–378. Ensinger, C., 2003. Her2/neu overexpression in differentiated thyroid carcino-
Jo, Y.S., Lee, J.C., Li, S., Choi, Y.S., Bai, Y.S., Kim, Y.J., Lee, I.S., Rha, S.Y., Ro, H.K., Kim, mas predicts metastatic disease. Virchows Arch. 442 (4), 322–328.
J.M., Shong, M., 2008. Significance of the expression of major histocompatibility Kushchayeva, Y., Duh, Q.Y., Kebebew, E., D’Avanzo, A., Clark, O.H., 2008. Comparison
complex class II antigen, HLA-DR and -DQ, with recurrence of papillary thyroid of clinical characteristics at diagnosis and during follow-up in 118 patients with
cancer. Int. J. Cancer 122 (4), 785–790. Hurthle cell or follicular thyroid cancer. Am. J. Surg. 195 (4), 457–462.
Jo, Y.S., Li, S., Song, J.H., Kwon, K.H., Lee, J.C., Rha, S.Y., Lee, H.J., Sul, J.Y., Kweon, G.R., Lang, B.H., Lo, C.Y., Chan, W.F., Lam, K.Y., Wan, K.Y., 2007a. Staging systems for pap-
Ro, H.K., Kim, J.M., Shong, M., 2006. Influence of the BRAF V600E mutation on illary thyroid carcinoma: a review and comparison. Ann. Surg. 245 (3), 366–
expression of vascular endothelial growth factor in papillary thyroid cancer. J. 378.
Clin. Endocrinol. Metab. 91 (9), 3667–3670. Lang, B.H., Lo, C.Y., Chan, W.F., Lam, K.Y., Wan, K.Y., 2007b. Staging systems for fol-
Johansson, P., Pavey, S., Hayward, N., 2007. Confirmation of a BRAF mutation- licular thyroid carcinoma: application to 171 consecutive patients treated in a
associated gene expression signature in melanoma. Pigment Cell Res. 20 (3), tertiary referral centre. Endocr. Relat. Cancer 14 (1), 29–42.
216–221. Lange, D., Sporny, S., Sygut, J., Kulig, A., Jarzab, M., Kula, D., Jarzab, B., 2006.
Jonklaas, J., Sarlis, N.J., Litofsky, D., Ain, K.B., Bigos, S.T., Brierley, J.D., Cooper, D.S., Hau- Histopathological diagnosis of thyroid cancer in a multicenter trial. Endokrynol.
gen, B.R., Ladenson, P.W., Magner, J., Robbins, J., Ross, D.S., Skarulis, M., Maxon, Pol. 57 (4), 336–342.
26 D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28

Lazar, V., Bidart, J.M., Caillou, B., Mahe, C., Lacroix, L., Filetti, S., Schlumberger, M., Mitsutake, N., Knauf, J.A., Mitsutake, S., Mesa Jr., C., Zhang, L., Fagin, J.A., 2005.
1999. Expression of the Na+ /I− symporter gene in human thyroid tumors: a Conditional BRAFV600E expression induces DNA synthesis, apoptosis, dediffer-
comparison study with other thyroid-specific genes. J. Clin. Endocrinol. Metab. entiation, and chromosomal instability in thyroid PCCL3 cells. Cancer Res. 65
84 (9), 3228–3234. (6), 2465–2473.
Lee, J.H., Lee, E.S., Kim, Y.S., 2007. Clinicopathologic significance of BRAF V600E muta- Mitsutake, N, Miyagishi, M., Mitsutake, S., Akeno, N., Mesa Jr., C., Knauf, J.A.,
tion in papillary carcinomas of the thyroid: a meta-analysis. Cancer 110 (1), Zhang, L., Taira, K., Fagin, J.A., 2006. BRAF mediates RET/PTC-induced mitogen-
38–46. activated protein kinase activation in thyroid cells: functional support for
Lee, X., Gao, M., Ji, Y., Yu, Y., Feng, Y., Li, Y., Zhang, Y., Cheng, W., Zhao, W., 2009. Anal- requirement of the RET/PTC–RAS–BRAF pathway in papillary thyroid carcino-
ysis of differential BRAF(V600E) mutational status in high aggressive papillary genesis. Endocrinology 147 (2), 1014–1019.
thyroid microcarcinoma. Ann. Surg. Oncol. 16 (2), 240–245. Modi, J., Patel, A., Terrell, R., Tuttle, R.M., Francis, G.L., 2003. Papillary thyroid car-
Liu, D., Hu, S., Hou, P., Jiang, D., Condouris, S., Xing, M., 2007a. Suppression of cinomas from young adults and children contain a mixture of lymphocytes. J.
BRAF/MEK/MAP kinase pathway restores expression of iodide-metabolizing Clin. Endocrinol. Metab. 88 (9), 4418–4425.
genes in thyroid cells expressing the V600E BRAF mutant. Clin. Cancer Res. 13 Musholt, T.J., Brehm, C., Hanack, J., von, W.R., Musholt, P.B., 2006. Identification of
(4), 1341–1349. differentially expressed genes in papillary thyroid carcinomas with and without
Liu, D., Liu, Z., Condouris, S., Xing, M., 2007b. BRAF V600E maintains proliferation, rearrangements of the tyrosine kinase receptors RET and/or NTRK1. J. Surg. Res.
transformation, and tumorigenicity of BRAF-mutant papillary thyroid cancer 131 (1), 15–25.
cells. J. Clin. Endocrinol. Metab. 92 (6), 2264–2271. Nakayama, H., Yoshida, A., Nakamura, Y., Hayashi, H., Miyagi, Y., Wada, N., Rino, Y.,
Liu, Z., Hou, P., Ji, M., Guan, H., Studeman, K., Jensen, K., Vasko, V., El-Naggar, A.K., Xing, Masuda, M., Imada, T., 2007. Clinical significance of BRAF (V600E) mutation and
M., 2008. Highly prevalent genetic alterations in receptor tyrosine kinases and Ki-67 labeling index in papillary thyroid carcinomas. Anticancer Res. 27 (5B),
phosphatidylinositol 3-kinase/akt and mitogen-activated protein kinase path- 3645–3649.
ways in anaplastic and follicular thyroid cancers. J. Clin. Endocrinol. Metab. 93 Ngan, E.S., Lang, B.H., Liu, T., Shum, C.K., So, M.T., Lau, D.K., Leon, T.Y., Cherny, S.S.,
(8), 3106–3116. Tsai, S.Y., Lo, C.Y., Khoo, U.S., Tam, P.K., Garcia-Barcelo, M.M., 2009. A germline
LiVolsi, V.A., Asa, S.L., 1994. The demise of follicular carcinoma of the thyroid gland. mutation (A339V) in thyroid transcription factor-1 (TITF-1/NKX2.1) in patients
Thyroid 4 (2), 233–236. with multinodular goiter and papillary thyroid carcinoma. J. Natl. Cancer Inst.
LiVolsi, V.A., Baloch, Z.W., 2004. Follicular neoplasms of the thyroid: view, biases, 101 (3), 162–175.
and experiences. Adv. Anat. Pathol. 11 (6), 279–287. Nikiforov, Y.E., 2002. RET/PTC rearrangement in thyroid tumors. Endocr. Pathol. 13
Loh, K.C, Greenspan, F.S., Dong, F., Miller, T.R., Yeo, P.P., 1999. Influence of lympho- (1), 3–16.
cytic thyroiditis on the prognostic outcome of patients with papillary thyroid Nikiforova, M.N., Caudill, C.M., Biddinger, P., Nikiforov, Y.E., 2002. Prevalence of
carcinoma. J. Clin. Endocrinol. Metab. 84 (2), 458–463. RET/PTC rearrangements in Hashimoto’s thyroiditis and papillary thyroid carci-
Lupi, C., Giannini, R., Ugolini, C., Proietti, A., Berti, P., Minuto, M., Materazzi, G., Elisei, nomas. Int. J. Surg. Pathol. 10 (1), 15–22.
R., Santoro, M., Miccoli, P., Basolo, F., 2007. Association of BRAF V600E mutation Nikiforova, M.N., Ciampi, R., Salvatore, G., Santoro, M., Gandhi, M., Knauf, J.A.,
with poor clinicopathological outcomes in 500 consecutive cases of papillary Thomas, G.A., Jeremiah, S., Bogdanova, T.I., Tronko, M.D., Fagin, J.A., Nikiforov,
thyroid carcinoma. J. Clin. Endocrinol. Metab. 92 (11), 4085–4090. Y.E., 2004. Low prevalence of BRAF mutations in radiation-induced thyroid
Malchoff, C.D., Malchoff, D.M., 2006. Familial nonmedullary thyroid carcinoma. Can- tumors in contrast to sporadic papillary carcinomas. Cancer Lett. 209 (1), 1–6.
cer Control 13 (2), 106–110. Nikiforova, M.N., Kimura, E.T., Gandhi, M., Biddinger, P.W., Knauf, J.A., Basolo, F.,
Marques, A.R., Espadinha, C., Frias, M.J., Roque, L., Catarino, A.L., Sobrinho, L.G., Zhu, Z., Giannini, R., Salvatore, G., Fusco, A., Santoro, M., Fagin, J.A., Nikiforov,
Leite, V., 2004. Underexpression of peroxisome proliferator-activated receptor Y.E., 2003a. BRAF mutations in thyroid tumors are restricted to papillary carci-
(PPAR)gamma in PAX8/PPARgamma-negative thyroid tumours. Br. J. Cancer 91 nomas and anaplastic or poorly differentiated carcinomas arising from papillary
(4), 732–738. carcinomas. J. Clin. Endocrinol. Metab. 88 (11), 5399–5404.
Masago, K., Asato, R., Fujita, S., Hirano, S., Tamura, Y., Kanda, T., Mio, T., Katakami, Nikiforova, M.N., Lynch, R.A., Biddinger, P.W., Alexander, E.K., Dorn, G.W., Tallini,
N., Mishima, M., Ito, J., 2009. Epidermal growth factor receptor gene mutations G., Kroll, T.G., Nikiforov, Y.E., 2003b. RAS point mutations and PAX8–PPAR
in papillary thyroid carcinoma. Int. J. Cancer 124 (11), 2744–2749. gamma rearrangement in thyroid tumors: evidence for distinct molecular path-
Mazzaferri, E.L., Jhiang, S.M., 1994. Long-term impact of initial surgical and medical ways in thyroid follicular carcinoma. J. Clin. Endocrinol. Metab. 88 (5), 2318–
therapy on papillary and follicular thyroid cancer. Am. J. Med. 97 (5), 418–428. 2326.
Mazzaferri, E.L., Kloos, R.T., 2001. Clinical review 128: current approaches to primary Nikiforova, M.N., Stringer, J.R., Blough, R., Medvedovic, M., Fagin, J.A., Nikiforov,
therapy for papillary and follicular thyroid cancer. J. Clin. Endocrinol. Metab. 86 Y.E., 2000. Proximity of chromosomal loci that participate in radiation-induced
(4), 1447–1463. rearrangements in human cells. Science 290 (5489), 138–141.
Mazzaferri, E.L., Young, R.L., 1981. Papillary thyroid carcinoma: a 10 year follow-up Oler, G., Camacho, C.P., Hojaij, F.C., Michaluart Jr., P., Riggins, G.J., Cerutti, J.M., 2008.
report of the impact of therapy in 576 patients. Am. J. Med. 70 (3), 511–518. Gene expression profiling of papillary thyroid carcinoma identifies transcripts
McCall, K.D., Harii, N., Lewis, C.J., Malgor, R., Kim, W.B., Saji, M., Kohn, A.D., Moon, correlated with BRAF mutational status and lymph node metastasis. Clin. Cancer
R.T., Kohn, L.D., 2007. High basal levels of functional toll-like receptor 3 (TLR3) Res. 14 (15), 4735–4742.
and noncanonical Wnt5a are expressed in papillary thyroid cancer and are coor- Oler, G., Cerutti, J.M., 2009. High prevalence of BRAF mutation in a Brazilian cohort
dinately decreased by phenylmethimazole together with cell proliferation and of patients with sporadic papillary thyroid carcinomas: correlation with more
migration. Endocrinology 148 (9), 4226–4237. aggressive phenotype and decreased expression of iodide-metabolizing genes.
McConahey, W.M., Hay, I.D., Woolner, L.B., van Heerden, J.A., Taylor, W.F., 1986. Cancer 115 (5), 972–980.
Papillary thyroid cancer treated at the Mayo Clinic. 1946 through 1970: initial Oler, G., Ebina, K.N., Michaluart Jr., P., Kimura, E.T., Cerutti, J., 2005. Investigation of
manifestations, pathologic findings, therapy, and outcome. Mayo Clin. Proc. 61 BRAF mutation in a series of papillary thyroid carcinoma and matched-lymph
(12), 978–996. node metastasis reveals a new mutation in metastasis. Clin. Endocrinol. (Oxf.)
McKay, J.D., Lesueur, F., Jonard, L., Pastore, A., Williamson, J., Hoffman, L., Burgess, 62 (4), 509–511.
J., Duffield, A., Papotti, M., Stark, M., Sobol, H., Maes, B., Murat, A., Kaariainen, Omidfar, K., Moinfar, Z., Sohi, A.N., Tavangar, S.M., Haghpanah, V., Heshmat, R.,
H., Bertholon-Gregoire, M., Zini, M., Rossing, M.A., Toubert, M.E., Bonichon, F., Kashanian, S., Larijani, B., 2009. Expression of EGFRvIII in thyroid carcinoma:
Cavarec, M., Bernard, A.M., Boneu, A., Leprat, F., Haas, O., Lasset, C., Schlumberger, immunohistochemical study by camel antibodies. Immunol. Invest. 38 (2),
M., Canzian, F., Goldgar, D.E., Romeo, G., 2001. Localization of a susceptibility 165–180.
gene for familial nonmedullary thyroid carcinoma to chromosome 2q21. Am. J. Pacini, F., Schlumberger, M., Dralle, H., Elisei, R., Smit, J.W., Wiersinga, W., 2006.
Hum. Genet. 69 (2), 440–446. European consensus for the management of patients with differentiated thyroid
Meijerink, J.P., den Boer, M.L., Pieters, R., 2009. New genetic abnormalities and treat- carcinoma of the follicular epithelium. Eur. J. Endocrinol. 154 (6), 787–803.
ment response in acute lymphoblastic leukemia. Semin. Hematol. 46 (1), 16–23. Palona, I., Namba, H., Mitsutake, N., Starenki, D., Podtcheko, A., Sedliarou, I., Oht-
Melck, A., Masoudi, H., Griffith, O.L., Rajput, A., Wilkins, G., Bugis, S., Jones, S.J., Wise- suru, A., Saenko, V., Nagayama, Y., Umezawa, K., Yamashita, S., 2006. BRAFV600E
man, S.M., 2007. Cell cycle regulators show diagnostic and prognostic utility for promotes invasiveness of thyroid cancer cells through nuclear factor kappaB
differentiated thyroid cancer. Ann. Surg. Oncol. 14 (12), 3403–3411. activation. Endocrinology 147 (12), 5699–5707.
Melillo, R.M., Castellone, M.D., Guarino, V., De Falco, V., Cirafici, A.M., Salvatore, G., Patel, K.N., Maghami, E., Wreesmann, V.B., Shaha, A.R., Shah, J.P., Ghossein, R., Singh,
Caiazzo, F., Basolo, F., Giannini, R., Kruhoffer, M., Orntoft, T., Fusco, A., Santoro, M., B., 2005. MUC1 plays a role in tumor maintenance in aggressive thyroid carci-
2005. The RET/PTC–RAS–BRAF linear signaling cascade mediates the motile and nomas. Surgery 138 (6), 994–1001.
mitogenic phenotype of thyroid cancer cells. J.Clin. Invest. 115 (4), 1068–1081. Pasieka, J.L., Zedenius, J., Auer, G., Grimelius, L., Hoog, A., Lundell, G., Wallin, G.,
Mesa Jr., C., Mirza, M., Mitsutake, N., Sartor, M., Medvedovic, M., Tomlinson, C., Backdahl, M., 1992. Addition of nuclear DNA content to the AMES risk-group
Knauf, J.A., Weber, G.F., Fagin, J.A., 2006. Conditional activation of RET/PTC3 and classification for papillary thyroid cancer. Surgery 112 (6), 1154–1159.
BRAFV600E in thyroid cells is associated with gene expression profiles that pre- Passler, C., Prager, G., Scheuba, C., Kaserer, K., Zettinig, G., Niederle, B., 2003. Appli-
dict a preferential role of BRAF in extracellular matrix remodeling. Cancer Res. cation of staging systems for differentiated thyroid carcinoma in an endemic
66 (13), 6521–6529. goiter region with iodine substitution. Ann. Surg. 237 (2), 227–234.
Min, H.S., Choe, G., Kim, S.W., Park, Y.J., Park, d.J., Youn, Y.K., Park, S.H., Cho, B.Y., Passler, C., Scheuba, C., Prager, G., Kaczirek, K., Kaserer, K., Zettinig, G., Niederle, B.,
Park, S.Y., 2008. S100A4 expression is associated with lymph node metastasis in 2004. Prognostic factors of papillary and follicular thyroid cancer: differences in
papillary microcarcinoma of the thyroid. Mod. Pathol. 21 (6), 748–755. an iodine-replete endemic goiter region. Endocr. Relat. Cancer 11 (1), 131–139.
Mineo, R., Costantino, A., Frasca, F., Sciacca, L., Russo, S., Vigneri, R., Belfiore, A., 2004. Pesutic-Pisac, V., Punda, A., Gluncic, I., Bedekovic, V., Pranic-Kragic, A., Kunac, N.,
Activation of the hepatocyte growth factor (HGF)-Met system in papillary thy- 2008. Cyclin D1 and p27 expression as prognostic factor in papillary carcinoma
roid cancer: biological effects of HGF in thyroid cancer cells depend on Met of thyroid: association with clinicopathological parameters. Croat. Med. J. 49 (5),
expression levels. Endocrinology 145 (9), 4355–4365. 643–649.
D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28 27

Peyssonnaux, C., Eychene, A., 2001. The Raf/MEK/ERK pathway: new concepts of Sancho, M., Vieira, J.M., Casalou, C., Mesquita, M., Pereira, T., Cavaco, B.M., Dias,
activation. Biol. Cell 93 (1–2), 53–62. S., Leite, V., 2006. Expression and function of the chemokine receptor CCR7 in
Pfister, S., Janzarik, W.G., Remke, M., Ernst, A., Werft, W., Becker, N., Toedt, G., thyroid carcinomas. J. Endocrinol. 191 (1), 229–238.
Wittmann, A., Kratz, C., Olbrich, H., Ahmadi, R., Thieme, B., Joos, S., Radlwimmer, Santoro, M., Melillo, R.M., Carlomagno, F., Vecchio, G., Fusco, A., 2004. Minireview:
B., Kulozik, A., Pietsch, T., Herold-Mende, C., Gnekow, A., Reifenberger, G., Kor- RET: normal and abnormal functions. Endocrinology 145 (12), 5448–5451.
shunov, A., Scheurlen, W., Omran, H., Lichter, P., 2008. BRAF gene duplication Sarquis, M.S., Weber, F., Shen, L., Broelsch, C.E., Jhiang, S.M., Zedenius, J., Frilling,
constitutes a mechanism of MAPK pathway activation in low-grade astrocy- A., Eng, C., 2006. High frequency of loss of heterozygosity in imprinted, com-
tomas. J. Clin. Invest. 118 (5), 1739–1749. pared with nonimprinted, genomic regions in follicular thyroid carcinomas and
Pierotti, M.A., Bongarzone, I., Borello, M.G., Greco, A., Pilotti, S., Sozzi, G., 1996. Cyto- atypical adenomas. J. Clin. Endocrinol. Metab. 91 (1), 262–269.
genetics and molecular genetics of carcinomas arising from thyroid epithelial Schlumberger, M., Challeton, C., De Vathaire, F., Travagli, J.P., Gardet, P., Lumbroso,
follicular cells. Genes Chromosomes Cancer 16 (1), 1–14. J.D., Francese, C., Fontaine, F., Ricard, M., Parmentier, C., 1996. Radioactive iodine
Podnos, Y.D., Smith, D., Wagman, L.D., Ellenhorn, J.D., 2005. The implication of lymph treatment and external radiotherapy for lung and bone metastases from thyroid
node metastasis on survival in patients with well-differentiated thyroid cancer. carcinoma. J. Nucl. Med. 37 (4), 598–605.
Am. Surg. 71 (9), 731–734. Shaha, A.R., Loree, T.R., Shah, J.P., 1994. Intermediate-risk group for differentiated
Polanski, A., Polanska, J., Jarzab, M., Wiench, M., Jarzab, B., 2007. Application of carcinoma of thyroid. Surgery 116 (6), 1036–1040.
Bayesian networks for inferring cause–effect relations from gene expression Sherman, S.I., 2003. Thyroid carcinoma. Lancet 361 (9356), 501–511.
profiles of cancer versus normal cells. Math. Biosci. 209 (2), 528–546. Sherman, S.I., Brierley, J.D., Sperling, M., Ain, K.B., Bigos, S.T., Cooper, D.S., Haugen,
Prasad, M.L., Pellegata, N.S., Kloos, R.T., Barbacioru, C., Huang, Y., de la Chapelle, A., B.R., Ho, M., Klein, I., Ladenson, P.W., Robbins, J., Ross, D.S., Specker, B., Taylor, T.,
2004. CITED1 protein expression suggests Papillary Thyroid Carcinoma in high Maxon III, H.R., National Thyroid Cancer Treatment Cooperative Study Registry
throughput tissue microarray-based study. Thyroid 14 (3), 169–175. Group, 1998. Prospective multicenter study of thyroid carcinoma treatment:
Puxeddu, E., Moretti, S., 2007. Clinical prognosis in BRAF-mutated PTC. Arq Bras. initial analysis of staging and outcome. Cancer 83 (5), 1012–1021.
Endocrinol. Metabol. 51 (5), 736–747. Shi, Y.F., Zou, M.J., Schmidt, H., Juhasz, F., Stensky, V., Robb, D., Farid, N.R., 1991. High
Puxeddu, E., Moretti, S., Elisei, R., Romei, C., Pascucci, R., Martinelli, M., Marino, C., rates of ras codon 61 mutation in thyroid tumors in an iodide-deficient area.
Avenia, N., Rossi, E.D., Fadda, G., Cavaliere, A., Ribacchi, R., Falorni, A., Pontecorvi, Cancer Res. 51 (10), 2690–2693.
A., Pacini, F., Pinchera, A., Santeusanio, F., 2004. BRAF(V599E) mutation is the Siragusa, M., Zerilli, M., Iovino, F., Francipane, M.G., Lombardo, Y., Ricci-Vitiani, L.,
leading genetic event in adult sporadic papillary thyroid carcinomas. J. Clin. Di, G.G., Todaro, M., De Maria, R., Stassi, G., 2007. MUC1 oncoprotein promotes
Endocrinol. Metab. 89 (5), 2414–2420. refractoriness to chemotherapy in thyroid cancer cells. Cancer Res. 67 (11),
Puxeddu, E., Zhao, G., Stringer, J.R., Medvedovic, M., Moretti, S., Fagin, J.A., 2005. Char- 5522–5530.
acterization of novel non-clonal intrachromosomal rearrangements between Skubis-Zegadlo, J., Nikodemska, A., Przytula, E., Mikula, M., Bardadin, K., Ostrowski, J.,
the H4 and PTEN genes (H4/PTEN) in human thyroid cell lines and papillary Wenzel, B.E., Czarnocka, B., 2005. Expression of pendrin in benign and malignant
thyroid cancer specimens. Mutat. Res. 570 (1), 17–32. human thyroid tissues. Br. J. Cancer 93 (1), 144–151.
Ramirez, R., Hsu, D., Patel, A., Fenton, C., Dinauer, C., Tuttle, R.M., Francis, G.L., 2000. Smallridge, R.C., Marlow, L.A., Copland, J.A., 2009. Anaplastic thyroid cancer: molec-
Over-expression of hepatocyte growth factor/scatter factor (HGF/SF) and the ular pathogenesis and emerging therapies. Endocr. Relat. Cancer 16 (1), 17–44.
HGF/SF receptor (cMET) are associated with a high risk of metastasis and recur- Smyth, P., Finn, S., Cahill, S., O’Regan, E., Flavin, R., O’Leary, J.J., Sheils, O., 2005. ret/PTC
rence for children and young adults with papillary thyroid carcinoma. Clin. and BRAF act as distinct molecular, time-dependant triggers in a sporadic Irish
Endocrinol. (Oxf.) 53 (5), 635–644. cohort of papillary thyroid carcinoma. Int. J. Surg. Pathol. 13 (1), 1–8.
Reddi, H.V., McIver, B., Grebe, S.K., Eberhardt, N.L., 2007. The paired box- Sorlie, T., Perou, C.M., Tibshirani, R., Aas, T., Geisler, S., Johnsen, H., Hastie, T., Eisen,
8/peroxisome proliferator-activated receptor-gamma oncogene in thyroid M.B., van de Rijn, M., Jeffrey, S.S., Thorsen, T., Quist, H., Matese, J.C., Brown, P.O.,
tumorigenesis. Endocrinology 148 (3), 932–935. Botstein, D., Eystein, L.P., Borresen-Dale, A.L., 2001. Gene expression patterns of
Rhoden, K.J., Unger, K., Salvatore, G., Yilmaz, Y., Vovk, V., Chiappetta, G., Qumsiyeh, breast carcinomas distinguish tumor subclasses with clinical implications. Proc.
M.B., Rothstein, J.L., Fusco, A., Santoro, M., Zitzelsberger, H., Tallini, G., 2006. Natl. Acad. Sci. U.S.A. 98 (19), 10869–10874.
RET/papillary thyroid cancer rearrangement in nonneoplastic thyrocytes: fol- Sturgis, C.D., Caraway, N.P., Johnston, D.A., Sherman, S.I., Kidd, L., Katz, R.L., 1999.
licular cells of Hashimoto’s thyroiditis share low-level recombination events Image analysis of papillary thyroid carcinoma fine-needle aspirates: significant
with a subset of papillary carcinoma. J. Clin. Endocrinol. Metab. 91 (6), 2414– association between aneuploidy and death from disease. Cancer 87 (3), 155–
2423. 160.
Riesco-Eizaguirre, G., Gutierrez-Martinez, P., Garcia-Cabezas, M.A., Nistal, M., San- Suarez, H.G., du Villard, J.A., Severino, M., Caillou, B., Schlumberger, M., Tubiana, M.,
tisteban, P., 2006. The oncogene BRAF V600E is associated with a high risk Parmentier, C., Monier, R., 1990. Presence of mutations in all three ras genes in
of recurrence and less differentiated papillary thyroid carcinoma due to the human thyroid tumors. Oncogene 5 (4), 565–570.
impairment of Na+ /I− targeting to the membrane. Endocr. Relat. Cancer 13 (1), Sugg, S.L., Ezzat, S., Rosen, I.B., Freeman, J.L., Asa, S.L., 1998. Distinct multiple
257–269. RET/PTC gene rearrangements in multifocal papillary thyroid neoplasia. J. Clin.
Roccato, E., Bressan, P., Sabatella, G., Rumio, C., Vizzotto, L., Pierotti, M.A., Greco, A., Endocrinol. Metab. 83 (11), 4116–4122.
2005. Proximity of TPR and NTRK1 rearranging loci in human thyrocytes. Cancer Takano, T., Ito, Y., Hirokawa, M., Yoshida, H., Miyauchi, A., 2007. BRAF V600E muta-
Res. 65 (7), 2572–2576. tion in anaplastic thyroid carcinomas and their accompanying differentiated
Rocha, A.S., Soares, P., Fonseca, E., Cameselle-Teijeiro, J., Oliveira, M.C., Sobrinho- carcinomas. Br. J. Cancer 96 (10), 1549–1553.
Simoes, M., 2003. E-cadherin loss rather than beta-catenin alterations is a Takano, T., Miyauchi, A., Yoshida, H., Kuma, K., Amino, N., 2004. High-throughput
common feature of poorly differentiated thyroid carcinomas. Histopathology differential screening of mRNAs by serial analysis of gene expression: decreased
42 (6), 580–587. expression of trefoil factor 3 mRNA in thyroid follicular carcinomas. Br. J. Cancer
Rodrigues, R., Roque, L., Espadinha, C., Pinto, A., Domingues, R., Dinis, J., Catarino, A., 90 (8), 1600–1605.
Pereira, T., Leite, V., 2007. Comparative genomic hybridization, BRAF, RAS, RET, Takayama, T., Miyanishi, K., Hayashi, T., Sato, Y., Niitsu, Y., 2006. Colorectal cancer:
and oligo-array analysis in aneuploid papillary thyroid carcinomas. Oncol. Rep. genetics of development and metastasis. J. Gastroenterol. 41 (3), 185–192.
18 (4), 917–926. Tallini, G., Asa, S.L., 2001. RET oncogene activation in papillary thyroid carcinoma.
Roque, L, Nunes, V.M., Ribeiro, C., Martins, C., Soares, J., 2001. Karyotypic character- Adv. Anat. Pathol. 8 (6), 345–354.
ization of papillary thyroid carcinomas. Cancer 92 (10), 2529–2538. Tanaka, K., Kurebayashi, J., Sonoo, H., Otsuki, T., Yamamoto, Y., Ohkubo, S.,
Rosenbaum, E., Hosler, G., Zahurak, M., Cohen, Y., Sidransky, D., Westra, W.H., 2005. Yamamoto, S., Shimozuma, K., 2002. Expression of vascular endothelial growth
Mutational activation of BRAF is not a major event in sporadic childhood papil- factor family messenger RNA in diseased thyroid tissues. Surg. Today 32 (9),
lary thyroid carcinoma. Mod. Pathol. 18 (7), 898–902. 761–768.
Ruan, D.T., Warren, R.S., Moalem, J., Chung, K.W., Griffin, A.C., Shen, W., Duh, Q.Y., Teixeira, M.R., 2006. Recurrent fusion oncogenes in carcinomas. Crit. Rev. Oncog. 12
Nakakura, E., Donner, D.B., Khanafshar, E., Weng, J., Clark, O.H., Kebebew, E., (3–4), 257–271.
2008. Mitogen-inducible gene-6 expression correlates with survival and is an Thompson, L.D., Wieneke, J.A., Paal, E., Frommelt, R.A., Adair, C.F., Heffess, C.S., 2001.
independent predictor of recurrence in BRAF(V600E) positive papillary thyroid A clinicopathologic study of minimally invasive follicular carcinoma of the thy-
cancers. Surgery 144 (6), 908–913. roid gland with a review of the English literature. Cancer 91 (3), 505–524.
Sahin, M., Allard, B.L., Yates, M., Powell, J.G., Wang, X.L., Hay, I.D., Zhao, Y., Goellner, Tian, X., Cong, M., Zhou, W., Zhu, J., Liu, Q., 2008. Relationship between protein
J.R., Sebo, T.J., Grebe, S.K., Eberhardt, N.L., McIver, B., 2005. PPARgamma staining expression of VEGF-C, MMP-2 and lymph node metastasis in papillary thyroid
as a surrogate for PAX8/PPARgamma fusion oncogene expression in follicu- cancer. J. Int. Med. Res. 36 (4), 699–703.
lar neoplasms: clinicopathological correlation and histopathological diagnostic Tomlins, S.A., Bjartell, A., Chinnaiyan, A.M., Jenster, G., Nam, R.K., Rubin, M.A.,
value. J. Clin. Endocrinol. Metab. 90 (1), 463–468. Schalken, J.A., 2009. ETS gene fusions in prostate cancer: from discovery to daily
Sakorafas, G.H., Sampanis, D., Safioleas, M., 2009. Cervical lymph node dissection in clinical practice. Eur. Urol. 56 (2), 275–286.
papillary thyroid cancer: current trends, persisting controversies, and unclari- Tubiana, M., Schlumberger, M., Rougier, P., Laplanche, A., Benhamou, E., Gardet,
fied uncertainties. Surg. Oncol., doi:10.1016/j.physletb.2003.10.071. P., Caillou, B., Travagli, J.P., Parmentier, C., 1985. Long-term results and prog-
Saltman, B., Singh, B., Hedvat, C.V., Wreesmann, V.B., Ghossein, R., 2006. Patterns nostic factors in patients with differentiated thyroid carcinoma. Cancer 55 (4),
of expression of cell cycle/apoptosis genes along the spectrum of thyroid carci- 794–804.
noma progression. Surgery 140 (6), 899–905. Tuttle, R.M., Leboeuf, R., Shaha, A.R., 2008. Medical management of thyroid cancer:
Sampson, E., Brierley, J.D., Le, L.W., Rotstein, L., Tsang, R.W., 2007. Clinical man- a risk adapted approach. J. Surg. Oncol. 97 (8), 712–716.
agement and outcome of papillary and follicular (differentiated) thyroid Ugolini, C., Giannini, R., Lupi, C., Salvatore, G., Miccoli, P., Proietti, A., Elisei, R., Santoro,
cancer presenting with distant metastasis at diagnosis. Cancer 110 (7), 1451– M., Basolo, F., 2007. Presence of BRAF V600E in very early stages of papillary
1456. thyroid carcinoma. Thyroid 17 (5), 381–388.
28 D. Handkiewicz-Junak et al. / Molecular and Cellular Endocrinology 322 (2010) 8–28

Unger, K., Zitzelsberger, H., Salvatore, G., Santoro, M., Bogdanova, T., Braselmann, H., ARHI contributes to follicular thyroid carcinogenesis. J. Clin. Endocrinol. Metab.
Kastner, P., Zurnadzhy, L., Tronko, N., Hutzler, P., Thomas, G., 2004. Heterogeneity 90 (2), 1149–1155.
in the distribution of RET/PTC rearrangements within individual post-Chernobyl Weber, F., Shen, L., Aldred, M.A., Morrison, C.D., Frilling, A., Saji, M., Schuppert, F.,
papillary thyroid carcinomas. J. Clin. Endocrinol. Metab. 89 (9), 4272–4279. Broelsch, C.E., Ringel, M.D., Eng, C., 2005b. Genetic classification of benign and
van’t Veer, L.J., Paik, S., Hayes, D.F., 2005. Gene expression profiling of breast cancer: malignant thyroid follicular neoplasia based on a three-gene combination. J. Clin.
a new tumor marker. J. Clin. Oncol. 23 (8), 1631–1635. Endocrinol. Metab. 90 (5), 2512–2521.
Vasko, V., Espinosa, A.V., Scouten, W., He, H., Auer, H., Liyanarachchi, S., Larin, Wiseman, S.M., Griffith, O.L., Deen, S., Rajput, A., Masoudi, H., Gilks, B., Goldstein,
A., Savchenko, V., Francis, G.L., de la Chapelle, A., Saji, M., Ringel, M.D., 2007. L., Gown, A., Jones, S.J., 2007. Identification of molecular markers altered during
Gene expression and functional evidence of epithelial-to-mesenchymal transi- transformation of differentiated into anaplastic thyroid carcinoma. Arch. Surg.
tion in papillary thyroid carcinoma invasion. Proc. Natl. Acad. Sci. U.S.A. 104 (8), 142 (8), 717–727.
2803–2808. Wiseman, S.M., Griffith, O.L., Melck, A., Masoudi, H., Gown, A., Nabi, I.R., Jones, S.J.,
Vasko, V., Ferrand, M., Di, C.J., Carayon, P., Henry, J.F., de Micco, C., 2003. Specific pat- 2008a. Evaluation of type 1 growth factor receptor family expression in benign
tern of RAS oncogene mutations in follicular thyroid tumors. J. Clin. Endocrinol. and malignant thyroid lesions. Am. J. Surg. 195 (5), 667–673.
Metab. 88 (6), 2745–2752. Wiseman, S.M., Melck, A., Masoudi, H., Ghaidi, F., Goldstein, L., Gown, A., Jones,
Vasko, V., Hu, S., Wu, G., Xing, J.C., Larin, A., Savchenko, V., Trink, B., Xing, M., 2005. S.J., Griffith, O.L., 2008b. Molecular phenotyping of thyroid tumors identifies
High prevalence and possible de novo formation of BRAF mutation in metasta- a marker panel for differentiated thyroid cancer diagnosis. Ann. Surg. Oncol. 15
sized papillary thyroid cancer in lymph nodes. J. Clin. Endocrinol. Metab. 90 (9), (10), 2811–2826.
5265–5269. Wreesmann, V.B., Sieczka, E.M., Socci, N.D., Hezel, M., Belbin, T.J., Childs, G., Patel,
Verburg, F.A., Mader, U., Luster, M., Reiners, C., 2009. Histology does not influ- S.G., Patel, K.N., Tallini, G., Prystowsky, M., Shaha, A.R., Kraus, D., Shah, J.P., Rao,
ence prognosis in differentiated thyroid carcinoma when accounting for age, P.H., Ghossein, R., Singh, B., 2004. Genome-wide profiling of papillary thyroid
tumour diameter, invasive growth and metastases. Eur. J. Endocrinol. 160 (4), cancer identifies MUC1 as an independent prognostic marker. Cancer Res. 64
619–624. (11), 3780–3789.
Viglietto, G, Chiappetta, G., Martinez-Tello, F.J., Fukunaga, F.H., Tallini, G., Wu, G., Mambo, E., Guo, Z., Hu, S., Huang, X., Gollin, S.M., Trink, B., Ladenson, P.W.,
Rigopoulou, D., Visconti, R., Mastro, A., Santoro, M., Fusco, A., 1995. RET/PTC Sidransky, D., Xing, M., 2005. Uncommon mutation, but common amplifica-
oncogene activation is an early event in thyroid carcinogenesis. Oncogene 11 tions, of the PIK3CA gene in thyroid tumors. J. Clin. Endocrinol. Metab. 90 (8),
(6), 1207–1210. 4688–4693.
von Rhein, W, Rhein, A., Werner, M., Scheumann, G.F., Dralle, H., Potter, E., Brabant, Xing, M., 2005. BRAF mutation in thyroid cancer. Endocr. Relat. Cancer 12 (2),
G., Georgii, A., 1997. Immunohistochemical detection of E-cadherin in differen- 245–262.
tiated thyroid carcinomas correlates with clinical outcome. Cancer Res. 57 (12), Xing, M., 2007. BRAF mutation in papillary thyroid cancer: pathogenic role, molec-
2501–2507. ular bases, and clinical implications. Endocr. Rev. 28 (7), 742–762.
Wagner, P.L., Moo, T.A., Arora, N., Liu, Y.F., Zarnegar, R., Scognamiglio, T., Fahey III, Xu, J., Moatamed, F., Caldwell, J.S., Walker, J.R., Kraiem, Z., Taki, K., Brent, G.A., Her-
T.J., 2008. The chemokine receptors CXCR4 and CCR7 are associated with tumor shman, J.M., 2003. Enhanced expression of nicotinamide N-methyltransferase
size and pathologic indicators of tumor aggressiveness in papillary thyroid car- in human papillary thyroid carcinoma cells. J. Clin. Endocrinol. Metab. 88 (10),
cinoma. Ann. Surg. Oncol. 15 (10), 2833–2841. 4990–4996.
Wajapeyee, N., Serra, R.W., Zhu, X., Mahalingam, M., Green, M.R., 2008. Oncogenic Yeh, J.J., Marsh, D.J., Zedenius, J., Dwight, T., Delbridge, L., Robinson, B.G., Eng, C.,
BRAF induces senescence and apoptosis through pathways mediated by the 1999. Fine-structure deletion mapping of 10q22–24 identifies regions of loss
secreted protein IGFBP7. Cell 132 (3), 363–374. of heterozygosity and suggests that sporadic follicular thyroid adenomas and
Wan, P.T., Garnett, M.J., Roe, S.M., Lee, S., Niculescu-Duvaz, D., Good, V.M., Jones, follicular thyroid carcinomas develop along distinct neoplastic pathways. Genes
C.M., Marshall, C.J., Springer, C.J., Barford, D., Marais, R., 2004. Mechanism of Chromosomes Cancer 26 (4), 322–328.
activation of the RAF–ERK signaling pathway by oncogenic mutations of B-RAF. Zaydfudim, V., Feurer, I.D., Griffin, M.R., Phay, J.E., 2008. The impact of lymph node
Cell 116 (6), 855–867. involvement on survival in patients with papillary and follicular thyroid carci-
Wang, Y., Hou, P., Yu, H., Wang, W., Ji, M., Zhao, S., Yan, S., Sun, X., Liu, D., Shi, B., noma. Surgery 144 (6), 1070–1077.
Zhu, G., Condouris, S., Xing, M., 2007. High prevalence and mutual exclusivity of Zhu, Z., Ciampi, R., Nikiforova, M.N., Gandhi, M., Nikiforov, Y.E., 2006. Prevalence of
genetic alterations in the phosphatidylinositol-3-kinase/akt pathway in thyroid RET/PTC rearrangements in thyroid papillary carcinomas: effects of the detec-
tumors. J. Clin. Endocrinol. Metab. 92 (6), 2387–2390. tion methods and genetic heterogeneity. J. Clin. Endocrinol. Metab. 91 (9),
Wang, Y., Ji, M., Wang, W., Miao, Z., Hou, P., Chen, X., Xu, F., Zhu, G., Sun, X., Li, Y., Con- 3603–3610.
douris, S., Liu, D., Yan, S., Pan, J., Xing, M., 2008a. Association of the T1799A BRAF Zhu, Z., Gandhi, M., Nikiforova, M.N., Fischer, A.H., Nikiforov, Y.E., 2003. Molecu-
mutation with tumor extrathyroidal invasion, higher peripheral platelet counts, lar profile and clinical–pathologic features of the follicular variant of papillary
and over-expression of platelet-derived growth factor-B in papillary thyroid thyroid carcinoma. An unusually high prevalence of ras mutations. Am. J. Clin.
cancer. Endocr. Relat. Cancer 15 (1), 183–190. Pathol. 120 (1), 71–77.
Wang, Y.L, Wang, J.C., Wu, Y., Zhang, L., Huang, C.P., Shen, Q., Zhu, Y.X., Li, D.S., Ji, Q.H., Zou, M., Al-Baradie, R.S., Al-Hindi, H., Farid, N.R., Shi, Y., 2005. S100A4 (Mts1) gene
2008b. Incidentally simultaneous occurrence of RET/PTC, H4-PTEN and BRAF overexpression is associated with invasion and metastasis of papillary thyroid
mutation in papillary thyroid carcinoma. Cancer Lett. 263 (1), 44–52. carcinoma. Br. J. Cancer 93 (11), 1277–1284.
Watanabe, R., Hayashi, Y., Sassa, M., Kikumori, T., Imai, T., Kiuchi, T., Murata, Y., Zou, M., Famulski, K.S., Parhar, R.S., Baitei, E., Al-Mohanna, F.A., Farid, N.R., Shi, Y.,
2009. Possible involvement of BRAFV600E in altered gene expression in papillary 2004. Microarray analysis of metastasis-associated gene expression profiling in
thyroid cancer. Endocr. J.. a murine model of thyroid carcinoma pulmonary metastasis: identification of
Weber, F., Aldred, M.A., Morrison, C.D., Plass, C., Frilling, A., Broelsch, C.E., Waite, S100A4 (Mts1) gene overexpression as a poor prognostic marker for thyroid
K.A., Eng, C., 2005a. Silencing of the maternally imprinted tumor suppressor carcinoma. J. Clin. Endocrinol. Metab. 89 (12), 6146–6154.

Das könnte Ihnen auch gefallen