Next Article in Journal
N-acetyltransferase Gene Variants Involved in Pediatric Idiosyncratic Drug-Induced Liver Injury
Next Article in Special Issue
The Role of Body Mass Index (BMI) in Differentiated Thyroid Cancer: A Potential Prognostic Factor?
Previous Article in Journal
First-Trimester Plasmatic microRNAs Are Associated with Fasting Glucose Levels in Late Second Trimester of Pregnancy
Previous Article in Special Issue
Molecular Pathology of Thyroid Tumors: Essential Points to Comprehend Regarding the Latest WHO Classification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Trends and Potential of Radiotherapy in the Treatment of Anaplastic Thyroid Cancer

1
Cancer Biology Division, Kanagawa Cancer Center Research Institute, Yokohama 2418515, Japan
2
Biospecimen Center, Kanagawa Cancer Center, Yokohama 2418515, Japan
3
Division of Cancer Immunotherapy, Kanagawa Cancer Center Research Institute, Yokohama 2418515, Japan
4
Department of Radiation Oncology, Kanagawa Cancer Center, Yokohama 2418515, Japan
5
Department of Breast and Thyroid Surgery, Yokohama City University Medical Center, Yokohama 2320024, Japan
6
Department of Charged Particle Therapy Research, QST Hospital, National Institutes for Quantum Science and Technology, Chiba 2638555, Japan
7
Department of Surgery, Yokohama City University, Yokohama 2360004, Japan
*
Author to whom correspondence should be addressed.
Biomedicines 2024, 12(6), 1286; https://doi.org/10.3390/biomedicines12061286
Submission received: 26 April 2024 / Revised: 4 June 2024 / Accepted: 7 June 2024 / Published: 10 June 2024

Abstract

:
Anaplastic thyroid cancer (ATC) is a rare but highly aggressive malignancy characterized by advanced disease at diagnosis and a poor prognosis. Despite multimodal therapeutic approaches that include surgery, radiotherapy, and chemotherapy, an optimal treatment strategy remains elusive. Current developments in targeted therapies and immunotherapy offer promising avenues for improved outcomes, particularly for BRAF-mutant patients. However, challenges remain regarding overcoming drug resistance and developing effective treatments for BRAF-wild-type tumors. This comprehensive review examines the clinical and biological features of ATC, outlines the current standards of care, and discusses recent developments with a focus on the evolving role of radiotherapy. Moreover, it emphasizes the necessity of a multidisciplinary approach and highlights the urgent need for further research to better understand ATC pathogenesis and identify new therapeutic targets. Collaborative efforts, including large-scale clinical trials, are essential for translating these findings into improved patient outcomes.

1. Introduction

Although anaplastic thyroid cancer (ATC) occurs rarely [1,2,3,4], it exhibits an extremely poor prognosis [5,6,7] and is characterized by the presence of unresectable local extension or distant metastasis at initial diagnosis [8,9]. A radical cure for this malignancy is extremely difficult. To maintain and prolong quality of life (QOL), a multidisciplinary approach combining surgery, radiotherapy, and chemotherapy is currently used, but this approach yields poor results. Recently, clinical trials with kinase inhibitors have been conducted in cases in which specific genetic mutations, such as BRAF V600E, have been identified, with the hope of improving the prognosis of ATC. One example of a successful study was the combination of the BRAF inhibitor dabrafenib and the MEK inhibitor trametinib in ATC patients harboring the BRAF V600E mutation [10]. Certain reports have suggested the possible beneficial role of immune checkpoint inhibitors [7,8,9]. However, molecular targeted agents can lead to the emergence of drug resistance during administration, and this is a significant factor that complicates cancer treatment. Additionally, patients who do not possess the target gene mutation are not eligible for these agents. Immunotherapy can cause serious side effects, such as interstitial pneumonia, due to an excessive immune response when T cells in the body are activated.
Here, we summarize the clinical and biological features of ATC and the current standard of care and its problems, and we discuss future prospects for the management of ATC. Several useful review articles exist regarding ATC management [11,12,13,14,15], with emphasis on molecular targeted agents. We would like to distinguish our review from these other manuscripts by highlighting the potential for ATC management in the context of radiotherapy based on the remarkable technological development that has occurred in recent years.

2. The Characteristics of ATC

2.1. Epidemiology and Clinical Presentation

Most thyroid cancers are differentiated thyroid cancers (DTC) originating from follicular cells, such as papillary thyroid cancer (PTC) and follicular thyroid cancer (FTC), generally exhibit a good prognosis [16]. In contrast, ATCs are extremely aggressive and exhibit a poor prognosis. They are considered to be one of the most malignant of all cancer types [17,18] and account for 10–35% of thyroid cancer deaths, despite their low incidence (approximately 1–2% of all thyroid cancers) [2,4,19]. The majority of patients with this disease are older than 60 years of age, and the sex ratio is the lowest compared to other types of thyroid cancer that are more common in women. Patients present with rapidly growing cervical mass and neck pain that develops within weeks, skin erythema, hoarseness, dysphagia, dyspnea, rapid swelling of the cervical lymph nodes, fever, fatigue, and weight loss. Invasion of the gastrointestinal tract is frequently observed. Patients may occasionally develop urgent airway narrowing, such as invasion or compression of the trachea, requires airway clearance to prevent death due to asphyxia. According to the eighth edition of the American Joint Committee on Cancer (AJCC)/Tumor-Node-Metastasis (TMN) cancer staging system [20], the disease is always in an advanced stage (stage IV) and is divided into IVA (localized stage) when the cancer is only within the thyroid gland, IVB (locally advanced stage) when there is gross extra-thyroidal extension or cervical lymph node metastasis, and IVC (metastatic stage) when there is distant metastasis [7,21] (Table 1).
Onoda et al. [22] evaluated overall survival (OS) using this system in the national database of the Thyroid Cancer Research Consortium for Japan (ATCCJ) [23], which consisted of 757 patients with ATC. The percentages and median OS for each stage included stage IVA (5.9%, 15.8 months), stage IVB (55.7%, 6.1 months), and stage IVC (38.3%, 2.8 months). Consequently, approximately 90% of patients with ATC exhibited disease progression to surrounding organs or distant metastases at initial diagnosis, and treatment may not have been initiated due to extremely rapid disease progression. Moreover, it is common for patients with ATC to die early after diagnosis if no effective treatment is available. Older patients may not tolerate aggressive radiotherapy or chemotherapy due to impaired immune and/or organ function or complications. Thus, age is a significant prognostic factor for ATC [24]. Additionally, leukocytosis (white blood cell count >10,000/mL), extrathyroidal extension, and distant metastases have been reported as prognostic factors [23,25].

2.2. Genomic Changes

Recently, several analytical studies using next-generation sequencing have been conducted to investigate the genetic alterations in ATC [6,26,27,28,29,30,31,32,33]. According to their findings, the most commonly identified mutations in ATC were TP53 and telomerase reverse transcriptase (TERT) promoter mutations. Although these mutations are observed in DTC and poorly differentiated DTC (PDTC), they are more frequent in ATC than they are in others [6,30]. TP53 is a well-known tumor suppressor gene involved in a variety of cellular functions, including cell survival, DNA repair, apoptosis, cell cycle checkpoints, and senescence. TP53 mutations promote cell proliferation and tumor progression through the loss of these functions [34]. This was strongly correlated with the malignant potential of ATC. Landa et al. [30] reported that TP53 mutations were observed in 8% of PDTC; however, it increased to 73% in ATC. The analyses by Pozdeyev et al. [28] and Xu et al. [6] revealed high frequencies of TP53 mutations (65% and 63%, respectively). TERT promoter mutations (C228T and C250T) are common in ATC. TERT is the catalytic subunit of telomerase and is essential for telomerase activity [35]. TERT is not expressed in most human somatic cells; however, its expression and transcription are upregulated in many cancers through various mechanisms, including mutations in the core promoter region of the TERT gene. Moreover, it elongates telomeres, confers unlimited proliferative capacity to tumor cells, and contributes to tumor progression and aggressiveness [36,37]. Recently, it has been reported that TERT interacts with various signaling molecules such as NF-κ B, c-MYC, β-catenin, and TCF-4 to increase cancer malignancy and promote cancer progression [38,39,40,41]. Landa et al. reported that TERT promoter mutations were observed at a high frequency of 73% in ATCs, compared to 9% in PTCs and 40% in PDTCs [30]. Xu et al. observed this mutation at a high frequency (75%) in ATCs [6].
Two major cascades are responsible for tumorigenesis in ATCs, including the mitogen-activated protein kinase (MAPK) signaling pathway (RAS-RAF-MEK-ERK) and the PI3K-Akt-mTOR (PAM) signaling pathway. As described in Figure 1, both pathways are involved in cell proliferation and survival and essentially start with receptor tyrosine kinase (RTK), which is activated by the binding of ligands such as growth factors such as fibroblast growth factor (FGF) or epidermal growth factor (EGF) to specific receptors. When one of these molecules is continuously activated by a genetic mutation, it consistently activates downstream molecules without signals from the upstream molecules.
The MAPK signaling pathway is involved in various cellular processes such as growth, proliferation, survival (avoidance of apoptosis [42] and induction of autophagy [43]), migration [44], and angiogenesis. The intracellular signaling of the MAPK pathway has been described in detail by Schubert et al. [45] and Cook et al. [46]. Common alterations in the MAPK pathway in ATC include mutations in BRAF and RAS. BRAF mutations are the primary therapeutic targets for this disease. The majority of BRAF mutations in ATC are BRAF V600E point mutations [29] that play an essential role in the development and progression of tumors [47,48]. This mutation activates BRAF kinase, which phosphorylates multiple targets, including MEK and ERK [49]. The BRAF V600E mutation is observed in approximately 20–40% of ATCs [50], and Jannin et al. [12]. reported that the frequency of BRAF mutations in ATCs appears to vary in each region. According to recent reports, BRAF mutations have been observed in 40–45% of ATC cases in the United States [6,30], 14–37% of cases in Europe [31,33], and 41–48% of cases in East Asia (Japan and Korea) [26,51]. Although the details are unclear, they suggest an association between ethnicity and iodine intake. BRAF mutations, the most common mutations originally observed in PTCs, strongly activate the MAPK pathway, ultimately leading to thyroid cell dedifferentiation [52]. Oishi et al. summarized the mutation profiles of PTCs and ATCs and observed that BRAF status and BRAF genotype matched between PTCs and ATCs in 18 of 21 cases [53]. Yoo et al. [26] and Xu et al. [6] reported similar results. Genomic associations have been demonstrated between PTCs and ATCs, thus suggesting that ATCs are derived from PTCs. RAS genes such as HRAS, KRAS, and NRAS have been described. RAS mutations are observed at a frequency of 10–50% in ATC, according to next generation sequencer (NGS) analysis data [27,28,54]. Among them, NRAS mutations are significantly more common in thyroid cancer [55]. Oishi et al. [53] reported HRAS and KRAS mutations at 5%, while NRAS mutations were detected at 18%. Lai et al. [56] confirmed a high frequency of NRAS (30%) compared to that of HRAS (0%) and KRAS (11%). RAS mutations promote the activation of the MAPK and PAM signaling pathways. KRAS mutants are major activators of the MAPK pathway, whereas NRAS mutants are activators of the PAM pathway [57]. These mutations were frequently detected in FTC [26], thus suggesting that a few ATC may have originated from FTC. Furthermore, certain studies have demonstrated that eukaryotic translation initiation factor 1A X-linked (EIF1AX) mutations often occur together with RAS mutations in ATCs and that a positive correlation between RAS and EIF1AX proteins results in increased expression of the oncogene c-MYC [30,58,59].
The PAM signaling pathway regulates various cellular processes, including metabolism, motility, proliferation, growth, and survival [60]. Intracellular signaling in the PAM pathway has been described in detail by Glaviano et al. [61] and Yang et al. [62]. Hyperactivation of the PAM pathway is observed in many cancers and contributes to accelerated cancer initiation and progression and the development of therapeutic resistance [63,64]. PAM promotes epithelial-mesenchymal transition (EMT) and metastasis [65,66]. Mutations in PIK3CA, PTEN, and AKT1 have been identified in the PAM pathway of ATC. PIK3CA, which encodes the p110α catalytic subunit, is frequently mutated in various cancers and is known to be associated with cell signaling, proliferation, invasion, and cancer development [67,68,69]. PIK3CA mutations were detected in less than 20% of ATC [29,31]; however, they were more frequent than they were in DTC or PDTC [28,30]. PTEN is a representative tumor suppressor gene similar to TP53 and is frequently mutated in various cancers. PTEN plays a critical role in regulating cell growth, proliferation, survival, migration, and invasion [70]. PTEN mutations have been demonstrated to activate the PAM pathway with a loss of function [71,72]. Additionally, loss of PTEN function in combination with alterations in TP53 has been reported to accelerate tumor progression [73]. PTEN mutations in ATCs were less than 20%. Their frequency was higher than that of PTCs and PDTCs and similar to that of FTCs [28,30]. AKT1 regulates key processes such as glucose metabolism, apoptosis, cell proliferation, transcription, and cell migration. However, the AKT1 gene mutation is rarer than that of PI3KCA and PTEN in ATC and accounts for less than 10% of mutations in most reports [28,31]. Although genetic mutations in AKT are rare, overactivation has been observed in many cancers, including ATC, resulting in tumorigenesis, growth, invasion, and drug resistance [74].
Other mutations were observed in Wnt-β-catenin pathway-related genes (CTNNB1, AXIN1, and APC) [75] and epigenetic-related genes such as SWI/SNF chromatin remodeling complex (ARID1A, SMARCB1, and PBRM1) and histone methyltransferases (KMT2A, KMT2C, KMT2D, and SETD2). Moreover, mutations in the cyclin-dependent kinase inhibitor 2A (CDKN2A) gene that encodes p16 have been observed in a few ATC cases [28,29]. Furthermore, anaplastic lymphoma kinase (ALK) mutations and fusions that activate both MAPK and PAM pathways have been detected at low rates [28,29]. Similarly, receptor tyrosine kinase (RET) fusions have been observed [76].
Although the details are discussed in Section 3.4, “Targeted Therapy,” many drugs specific to these targets are currently under development and in clinical trials. Among these, combination therapy with a BRAF inhibitor and MEK inhibition has exhibited favorable antitumor effects in patients with ATC and the BRAF V600E mutation [10].

3. Treatment of ATC

ATC is difficult to treat as it is aggressive, spreads rapidly within the neck, and possesses the potential to metastasize to distant body sites. After considering the available therapies, comorbidities, and patient wishes, physicians decide whether to provide aggressive treatment or palliative care based on staging and prognosis. In general, radical treatment is difficult to achieve, and multidisciplinary treatment combining surgery, radiotherapy, and chemotherapy is used to prolong life and maintain QOL, except for a few cases (Figure 2). In retrospective studies, patients who received this multimodal therapy exhibited better prognoses than those who did not [17,23,77,78,79]. Radioactive iodine (RAI) therapy and thyroid-stimulating hormone (TSH) suppression that are commonly used to treat DTC are not effective for ATC [7]. In recent years, targeted therapy has been used with a few successes in cases with specific genetic mutations. Additionally, immune checkpoint inhibitors have been used successfully in a few cases.
In this section, we describe each treatment modality’s current status and future perspectives, considering the results of representative recent studies (Table 2) and ongoing clinical trials (Table 3) of multimodal treatments.

3.1. Surgery

A complete surgical excision followed by adjuvant therapy is the best approach to curing ATC [15]. However, curative resection is currently beneficial in early stage cases (IVA and part of IVB), as it may lead to long-term survival [86], whereas surgery must be carefully approved to maintain QOL in cases where the long-term prognosis is not expected [87]. Specifically, R0/R1 resection can be expected if the tumor does not extend beyond the common carotid artery, whereas tumors that extend beyond the common carotid artery are often inoperable. R0 resection, also known as curative resection, indicates a microscopically margin-negative resection with no residual tumor grossly or microscopically in the primary tumor. R1 resection indicates that all macroscopic lesions are removed, but the microscopic margins are tumor-positive [88,89]. Debulking surgery that minimizes postoperative QOL is considered valuable [90,91]; however, the extent to which it affects prognosis remains unclear. The National Comprehensive Cancer Network recommends total thyroidectomy with therapeutic lymph node dissection (R0/R1) and local radiotherapy if resectable, and tracheostomy with steroids only if more strongly indicated [9,92]; However, as recommended by the American Thyroid Association, it may be practical to perform surgery for local control while taking care not to interfere with other available palliative approaches, including radiation and systemic therapy, and not to compromise QOL [7].
The role of surgery in the treatment of locally advanced and metastatic ATC is currently the subject of considerable debate, with successful cases reported in patients with ATC harboring BRAF V600E mutations who received preoperative dabrafenib plus trametinib and ultimately underwent surgery. Looking to the future, this suggests that the use of surgery may be re-evaluated, even in advanced stages.

3.2. Radiotherapy

Rapid local progression and recurrence of ATC are associated with the aggressive nature of the disease, and local control using surgery or external beam radiotherapy (EBRT) is important. EBRT is a treatment in which X-rays are delivered to the tumor from high-energy radiotherapy equipment located outside of the body. Radiotherapy is recommended, as EBRT is used as a preoperative or postoperative adjuvant therapy and has been demonstrated to improve the median OS in retrospective studies [18,54,93,94]. For example, in a retrospective analysis of 496 patients with ATC, Saeed et al. reported that the survival of 375 patients who received adjuvant EBRT was significantly longer than that of 121 patients who did not receive adjuvant EBRT (12.3 vs. 9.1 months) [95]. The total dose was an important factor in EBRT. Most studies have indicated that OS and local control can be predicted by the total dose [96] that is better managed with irradiation greater than 45 Gy. For example, in a National Cancer Database analysis of 1,288 patients with unresected stage IVB and IVC ATC, Pezzi et al. reported that 1-year OS rates were improved in patients treated with 60–75 Gy compared to those treated with less than 60 Gy (31% vs. 16%) [97]. In a retrospective analysis by Fan et al., delivering 60 Gy or more was significantly associated with a lower risk of local progression and longer OS (10.6 vs. 3.6 months) [80]. Although the conventional irradiation regimen of 2 Gy once daily, as recommended by the American Thyroid Association and National Comprehensive Cancer Network guidelines, has been used as the standard option, hyperfractionated, accelerated, and fractionated irradiation have been studied. Hyperfractionated and accelerated irradiation have been considered to overcome the rapid progression and radioresistance of ATC. While certain reports suggest improved local control [98,99,100], others conclude that it neither reduces toxicity nor improves outcomes [101,102], ultimately leading to controversy. In contrast, hypofractionated irradiation, which delivers a higher dose in a shorter time, is used to improve quality of life and local control. Several clinical trials have exhibited promising results using hypofractionated RT for the treatment of ATC [103,104]. However, as discussed by Oliinyk et al. [102], limited data are evaluating hypofractionated regimens, and their use in the actual treatment of ATC remains conclusive. Irradiation techniques play an important role. The thyroid gland is located near the spinal cord, and this makes it difficult to avoid the spinal cord and deliver high doses to localized areas of cancer. One technology that makes this possible is intensity-modulated radiation therapy (IMRT). IMRT is a technique that delivers a high dose to the tumor while minimizing the dose to organs at risk (OAR) by varying the intensity of the radiation in the field during treatment, thus allowing precise dose delivery even to tumors with complex shapes. IMRT can precisely deliver radiation to thyroid cancer cells while reducing the radiation dose to the spinal cord [105]. IMRT is recommended as the standard of care for radiation therapy in ATC due to its advantages in reducing local recurrence, toxicity, and treatment complications [7,8]. Recently, volumetric modulated arc therapy has been developed, in which the device is rotated at different speeds and dose rates to modulate intensity, ultimately resulting in shorter treatment times and improved treatment accuracy [106]. Altogether, it is currently accepted that the appropriate treatment is to utilize IMRT to achieve a total dose of 60 Gy or more at 2 Gy per dose. Radiotherapy is often used in combination with chemotherapy, molecular targeted therapy, and immunotherapy to treat ATC. The clinical trials of combination therapies, including radiotherapy, are shown in Table 4.
However, photon beam radiotherapy exhibits limitations regarding delivering a radical dose to the tumor site while accounting for damage to the OAR. From a future perspective, one way to overcome this problem is to use particle beam radiotherapy such as proton beam radiotherapy (PBRT) and carbon ion radiotherapy (CIRT). Particle beams enter the body with low energy when they enter the surface, transfer most of their energy to a certain depth that is called the Bragg peak, and then decay rapidly. These physical characteristics allow particle beams to deliver higher doses to cancers while protecting the OAR compared to radiation therapy using photon beams. The relative biological effectiveness (RBE) of a proton beam is 1.1- to 1.2-fold higher than that of X-rays [109], whereas that of carbon ion beams is 2- to 3-fold higher than that of X-rays [110]. As higher RBE values indicate greater cell-killing capacity, particle beams are more effective for cancer treatment than X-rays [111]. The only actual application of particle therapy for ATC was reported by Youssef et al. [112]. Patients with ATC with recurrence after thyroidectomy and radioiodine therapy were treated with two cycles of intensity-modulated PBRT (IMPT) with QUAD shot radiation (3.7 Gy delivered in four fractions twice daily at least 6 h apart for 2 consecutive days, repeated every 4 weeks, and with concurrent chemotherapy with doxorubicin and dacarbazine). At the last follow-up at 12 months, the disease had not progressed locally or systemically, thus indicating that PBRT was effective for treating ATC. Carbon ion beams are even more biologically effective than proton beams and thus exert more promising therapeutic effects. Recently, intensity-modulated CIRT (IMCT) was developed using pencil beam scanning technology, in which the total dose is made uniform by the sum of the individual beams, but the intensity of each beam is different as in IMRT [113]. IMCT can further reduce the dose of irradiation to neighboring OARs compared to that of conventional CIRT. Although not included in ATC, there have been several reports detailing the results of IMCT in head and neck cancer that indicate a favorable therapeutic effect and a low incidence of acute and late toxicity [114,115]. Multi-ion radiotherapy (MIRT) is currently under development. Moreover, it combines ion beams with different linear energy transfers (LETs) of noncarbon ions with carbon ion beams, such as higher LET beams (e.g., oxygen and neon ions) targeting areas of resistance and lower LET beams (e.g., helium ions) for the boundary with normal tissue near the tumor. This approach offers a promising new treatment option for patients with complex cancers [116]. Additionally, the higher LET beams used in MIRT are more effective against radioresistant hypoxic cells than are carbon ions, thus potentially leading to earlier tumor reoxygenation [117,118,119].

3.3. Chemotherapy

Anthracyclines such as doxorubicin, platinum-based drugs such as cisplatin and carboplatin, and taxanes such as paclitaxel and docetaxel are the main chemotherapeutic drugs used for ATC [81]. However, the effects of these agents on ATCs are moderate and transient and are often disappointing. These guidelines recommend the rapid use of paclitaxel + carboplatin, docetaxel + doxorubicin, paclitaxel alone, or docetaxel alone when there are no therapeutic targets without molecular abnormalities or when targeted therapies are not available [7,8]. They can be used as bridging chemotherapeutic agents before the use of molecularly targeted agents [82]. Among these, taxanes have been reported to be effective for the treatment of ATC. For example, Higashiyama et al. treated stage IVB and IVC patients with ATC and weekly PTX induction chemotherapy and observed complete (CR) and partial response (PR) in 8% and 23% of patients, respectively [120]. They reported that the OS rate of paclitaxel-treated patients with stage IVB disease was better than that of patients who did not receive chemotherapy or those who received drugs other than paclitaxel. However, it is challenging to treat ATC with anticancer drugs alone.
Therefore, a future strategy is to use them in combination with other modalities. For example, taxanes increase the percentage of radiosensitive G2/M phase cells; as a result, their combination with radiotherapy may be effective. A report evaluated the use of docetaxel in combination with radiotherapy in patients with ATC and observed that 67% of patients achieved CR and the remaining 33% achieved PR, although the number of patients was small [121]. They reported that 83% of patients were alive at a median follow-up of 21.5 months. These findings suggest that a combination of taxane-based chemotherapy and radiotherapy may be effective for patients with ATC.

3.4. Targeted Therapy

Recently, genome sequencing analysis revealed molecules that are frequently mutated in ATCs and are involved in tumor progression and malignant transformation. Therapeutic agents targeting these molecules have been investigated. These agents target receptor tyrosine kinases (RTK) and downstream molecules of the MAPK and PAM signaling pathways.

3.4.1. Inhibitors of MAPK Pathway

Inhibitors of the MAPK pathway are potential agents for cancer therapy, and many compounds have been identified in clinical and preclinical trials. Several RAF and MEK inhibitors have been approved; however, no ERK inhibitors have yet been approved.
Dabrafenib is a BRAFV600E kinase inhibitor, and trametinib is an MEK inhibitor. Their combination therapy has been approved by the U.S. Food and Drug Administration (FDA) for the treatment of patients with ATC and BRAF V600E mutations. This is due to resistance to BRAF inhibitors that develops within months of BRAF blockade alone [122]. Several mechanisms have been proposed to induce this resistance, including the expression of splicing variants of the BRAF gene [123], inhibition of the negative feedback mechanism in the MAPK pathway [124], and activation of integrin/FAK signaling [125]. In vitro studies revealed that dabrafenib and trametinib inhibit ATC cell proliferation [126]. Subbiah et al. reported that in a phase II study evaluating the efficacy of dabrafenib plus trametinib in 36 patients with ATC, the CR and PR were 8% and 47%, respectively, and the median OS was 14.5 months [10]. In a recently reported UK clinical trial, CR and PR were achieved in 12% and 71% of patients with ATC, respectively, with a median OS of 6.9 months [127].
Another combination of BRAF and MEK inhibitors includes encorafenib (BRAF inhibitor) and binimetinib (MEK inhibitor). In a phase II study of 22 patients with thyroid cancer, including five patients with ATC and BRAF mutation-positive ATC with local invasion or distant metastasis, that evaluated the efficacy of the combination of encorafenib and binimetinib, the overall response rate (ORR) was 80% (CR, one case; PR, three cases) [128].
Vemurafenib is a selective BRAF inhibitor approved by the FDA for the treatment of patients with metastatic melanoma harboring the BRAF V600E mutation [129]. Zhang et al. reported that vemurafenib inhibits tumor growth in an in vivo xenograft mouse model using an ATC cell line [130]. A clinical trial evaluating the efficacy of vemurafenib in patients with cancer due to the BRAF V600E mutation, including seven patients with ATC, indicated that the percentages of patients with cancer achieving CR and PR were 14% and 14%, respectively [131]. Clinical trials are currently evaluating the efficacy of combining vemurafenib with cobimetinib, a MEK inhibitor, in BRAF-positive thyroid cancers, including ATC.
Other drugs, such as PLX8394 (BRAF inhibitor) and selumetinib (MEK inhibitor), have been investigated for the treatment of thyroid cancer [132]. Additionally, the ERK inhibitor DEL-22379 has been reported to decrease cell viability and inhibit the metastasis of ATC cells with BRAF mutations in vitro and in vivo [133].

3.4.2. Inhibitors of PAM Pathway

Molecules involved in this signaling pathway have received considerable attention in recent years, and many drugs targeting them have been studied and evaluated in animals and humans.
Everolimus inhibits the mTOR complex 1 (mTORC1). Owonikoko et al. reported that everolimus is effective against ATC cells both in vitro and in vivo [134]. However, the results of clinical trials investigating the efficacy of everolimus have been disappointing [83,135,136]. mTOR exists in two functionally and structurally distinct complexes, mTOR complex 1 (mTORC1) and mTORC2. Inhibition of mTORC1 inhibits ribosome biogenesis and suppresses cell proliferation, whereas inhibition of mTORC2 activates AKT and promotes cell growth [137]. This may have been due to the limited efficacy of everolimus.
In contrast, ATP-competitive dual mTORC1/2 inhibitors such as MLN0128 [138], AZD2014 [139,140], and AZD8055 [141] have been developed that can efficiently block the PAM pathway without causing negative feedback induction of mTORC2. Among these compounds, a phase II clinical trial is currently underway to evaluate the efficacy of MLN0128 for the treatment of metastatic ATC.
Other compounds, including CUDC-907 (PI3K and HDAC inhibitor) [142], copanlisib (PI3K inhibitor) [143], and buparlisib (PI3K inhibitor) [144], have been studied in the context of thyroid cancer.

3.4.3. RTK Inhibitors

RTKs are transmembrane glycoproteins on the cell surface that regulate cell proliferation, differentiation, and survival by activating several important intracellular signaling pathways, such as the MAPK and PAM pathways. RTKs include vascular endothelial growth factor receptor (VEGFR), epidermal growth factor receptor (EGFR), fibroblast growth factor receptor (FGFR), cluster of differentiation 117 (c-KIT), rearranged during transfection (RET), and platelet-derived growth factor receptor (PDGFR). Overexpression and activation of mutations in cancer cells have been reported, and drugs targeting them have been developed.
Lenvatinib is an inhibitor of VEGFR, FGFR, c-kit, RET, and PDGFR, and it is approved by the FDA and the European Medicines Agency for the treatment of iodine-131 (I-131) refractory DTC [145]. Ferrari et al. reported that lenvatinib suppresses the proliferative capacity of ATC cells in vitro and in vivo [146]. In a clinical trial of patients with ATC, Takahashi et al. reported a median OS of 10.4 months and an ORR of 24% in the lenvatinib group [84]. Reports have suggested that lenvatinib provides clinical benefits to patients with advanced ATC [147,148]. Although a high response rate can be expected, there is a risk of a shortened prognosis in patients with tracheal or carotid artery invasion due to fistula formation. Furthermore, lenvatinib alone is not an effective treatment for ATC [149] and has been disappointing in terms of prolonging the survival of unresectable ATC [150].
Similar to lenvatinib, single-agent multi-kinase inhibitors (MKIs), including sorafenib, gefitinib, imatinib, sunitinib, and pazopanib, have exhibited efficacy in vitro and in vivo [151,152,153,154,155,156,157] but have not exhibited promising results in clinical trials [107,158,159,160,161,162,163]. These MKIs are not effective when prescribed alone for the treatment of ATC but may be potentially useful in combination with other targeted therapies. Furthermore, the therapeutic effects of TRK or RET inhibitors are expected to be effective when mutations or fusions of TRK or RET are observed.
Targeted drugs have been developed against various target molecules, and a few of them, such as dabrafenib plus trametinib and encorafenib plus binimetinib, have demonstrated therapeutic efficacy. However, the emergence of drug resistance within a few months after a successful response has become a problem. Overcoming drug-resistant cancer is a major challenge in the development of molecularly targeted cancer therapies. Currently, there are no concrete, effective means to solve this problem.
Future prospects include elucidating the mechanisms of drug resistance at the basic research level and, in clinical practice, identifying effective therapeutic agents by studying the mutations that occur when drug resistance develops through genetic testing.

3.5. Immunotherapy

Tumor cells grow and metastasize through various mechanisms to escape recognition and attack by the immune system. Tumor-induced immunosuppression occurs in two ways. The first occurs through the induction of immunosuppressive cells that accumulate around the tumor and secrete immunosuppressive factors. Another mechanism involves the expression of immunosuppressive molecules and their receptors, such as programmed death ligand/programmed death-1 (PD-L1/PD-1), galectin-9/TIM-3, LAG-3, and CTLA-4. These are known as immune checkpoints that inhibit the activation of effector T lymphocytes, ultimately leading to the immune escape of the tumor. A recent trend in immunotherapy is the inhibition of these immune checkpoints and the restoration of immune function.
PD-L1 expressed on tumor cells binds to PD-1 on T cells and suppresses the proliferation and function of T cells, thereby weakening their immune response to tumor cells. PD-L1 expression in cancer cells may be primarily expressed or induced in the cancer microenvironment by the stimulation of inflammatory cytokines produced by tumor-infiltrating lymphocytes (TILs) and tumor-associated macrophages (TAMs) surrounding tumor cells [164]. PD-L1 is highly expressed in ATC [85,165,166], and Cantara et al. reported that immunohistochemistry using two PD-L1 antibodies exhibited high positivity rates of 65% and 90%, respectively [167]. Therefore, anti-PD-L1 and anti-PD-1 antibodies may be effective therapies for ATC. In contrast, CTLA-4 expressed on T cells binds to B7 molecules (CD80 and CD86) on antigen-presenting cells and inhibits T cell activation. CTLA-4 is expressed in regulatory T cells (Tregs). When Tregs bind to antigen-presenting cells, T cells cannot bind to antigen-presenting cells and cannot be activated. Therefore, anti-CTLA-4 antibodies are expected to be used for cancer therapy.
There are reports of the use of the anti-PD-1 antibodies spartalizumab and pembrolizumab and the anti-PD-L1 antibody atezolizumab in the treatment of ATC. In a phase II study by Capdevila et al. involving 42 patients with ATC treated with spartalizumab intravenously at a dose of 400 mg every 4 weeks, the ORR was 19% (CR, 7%; PR, 12%), and it was 29% in patients with PD-L1 expression and even higher in patients with high PD-L1 expression. The median OS was 5.9 months, with 40% of patients surviving for 1 year [85]. In a retrospective study, Iyer et al. demonstrated the efficacy of pembrolizumab in combination with lenvatinib, trametinib, or a combination of dabrafenib and trametinib in patients with ATC [168]. Moreover, there were no CRs, but 42% of patients achieved OR. The median OS after the addition of pembrolizumab was 6.93 months. Wang et al. investigated six patients with ATC and BRAF V600E mutation who underwent complete surgical resection and received dabrafenib plus trametinib with or without pembrolizumab [169]. No local recurrence was detected in any patient. Four patients with pembrolizumab achieved CR within the observation period (7.8–26.0 months), while two patients without pembrolizumab died due to distant metastasis. These results suggest that the combination of dabrafenib and trametinib with ICI would be more effective for patients with ATC and the BRAF V600E mutation. A phase 2 pilot study is ongoing to evaluate the efficacy of dabrafenib and trametinib in combination with cemiplimab, an anti-PD-1 antibody, in the context of ATC treatment. Moreover, reports suggest that a combination of lenvatinib and pembrolizumab may be effective for treating ATC [170]. A study evaluating the efficacy of lenvatinib in combination with nivolumab is ongoing in Japan (jRCT2080224758), and results are expected in the near future. While immunotherapy is successful in a few cases, it can cause side effects. Immune checkpoint inhibitors reactivate immune cells that have been suppressed by cancer cells. These side effects are known as immune-related adverse events (irAEs) and have been reported to include interstitial pneumonia, colitis, type 1 diabetes, endocrine disorders such as thyroid dysfunction, liver and kidney dysfunction, skin disorders, myasthenia gravis, myositis, and uveitis [171,172]. Chronic side effects have been reported, such as those that persist for more than three months after the end of treatment with immune checkpoint inhibitors [173].
The irAEs are becoming very complicated due to the combination of immunotherapy and other therapies. In the future, it is necessary to move toward establishing and managing a safety management system across medical departments and professions in collaboration with multiple professions to address these issues.

3.6. Others

Recently, it has been increasingly recognized that microbiota, including those in the gut and oral cavity, influence the pathogenesis, prognosis, and treatment response of various diseases, including cancer [174]. Several studies have reported a relationship between thyroid cancer and the microbiota. Research on the relationship between oral bacteria has reported that an increase in the genera Alloprevotella, Anaeroglobus, and Acinetobacter, unclassified Bacteroidales, and unclassified Cyanobacteriales was observed in the saliva of patients with thyroid cancer [175,176]. Moreover, studies exist on the relationship between gut microbiota and thyroid cancer. The relative abundances of Neisseria and Streptococcus were significantly higher in the gut flora of patients with thyroid cancer than those in healthy controls [177]. A study examining patients with PTC revealed that the abundance of Firmicutes was higher in the fecal flora than it was in healthy participants [178].
Previously, the thyroid gland was believed to be a sterile organ; however, recent technological advances have revealed that it is colonized by microorganisms. Microorganisms exist in thyroid cancer tissues, and a relationship between bacterial flora and disease status has been reported. One study has demonstrated a correlation between Sphingomonas abundance within the tumor microenvironment and lymph node metastasis [179]. A study examining patients with PTC revealed a high abundance of Proteobacteria in the tumors. Additionally, tumor bacterial diversity increased in patients with a higher T stage [178]. Regarding the therapeutic response to postoperative I-131 treatment for PTC, low levels of the genera Dorea and Bifidobacterium were associated with a poor prognosis [180]. However, there have been no reports investigating the relationship between ATC and the bacterial flora. Treatment methods for ATC, such as radiotherapy, chemotherapy, and immunotherapy, are associated with bacterial flora [181]. Therefore, in the future, microflora will attract attention as a target for new treatment strategies for ATC, and further research and development are expected [182].

3.7. Best Supportive Care

Although promising therapies have been developed for the treatment of ATC in recent years, it remains a difficult-to-cure cancer with a high mortality rate. The patient wishes to be fully considered when deciding whether to pursue aggressive treatment, including palliative care for pain and dyspnea and hospice care [183].

4. Conclusions

Although the treatment of ATC is evolving and diversifying, the optimal treatment has not yet been identified, ultimately resulting in a poor prognosis. The rapid progression of ATC necessitates immediate diagnosis and treatment. A multidisciplinary care team, including endocrinologists, oncologic surgeons, medical oncologists, radiation oncologists, radiologists, and palliative care staff, must be formed to determine the treatment plan and provide appropriate care as soon as possible while considering patient wishes. Recent studies have demonstrated that BRAF-mutant ATCs exhibit dramatically improved prognosis. However, the problem of drug resistance and treatment strategies for wild-type BRAF remain underexplored. A better understanding of the molecular biology of ATCs is expected to lead to the development of novel targeted therapies. Although radiotherapy remains the mainstay of ATC treatment, recent technological developments such as intensity-modulated irradiation and particle therapy are remarkable. The successful use of combinations of these novel drugs and therapies may lead to the development of treatment strategies to improve the ATC prognosis. Further clinical trials using many patients are required.

Author Contributions

Conceptualization, K.S. and D.H.; methodology, K.S. and D.H.; data curation, K.S., H.H. and D.H.; writing—original draft preparation, K.S., H.H. and D.H.; writing—review and editing, K.S., H.H., S.T., N.S. (Nao Saito), R.H., N.S. (Nobuyasu Suganuma), T.S. and D.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef] [PubMed]
  2. Janz, T.A.; Neskey, D.M.; Nguyen, S.A.; Lentsch, E.J. Is the incidence of anaplastic thyroid cancer increasing: A population based epidemiology study. World J. Otorhinolaryngol. Head Neck Surg. 2019, 5, 34–40. [Google Scholar] [CrossRef] [PubMed]
  3. Casali, P.G.; Trama, A. Rationale of the rare cancer list: A consensus paper from the Joint Action on Rare Cancers (JARC) of the European Union (EU). ESMO Open 2020, 5, e000666. [Google Scholar] [CrossRef] [PubMed]
  4. Pereira, M.; Williams, V.L.; Hallanger Johnson, J.; Valderrabano, P. Thyroid Cancer Incidence Trends in the United States: Association with Changes in Professional Guideline Recommendations. Thyroid 2020, 30, 1132–1140. [Google Scholar] [CrossRef] [PubMed]
  5. Takahashi, S.; Tahara, M.; Ito, K.; Tori, M.; Kiyota, N.; Yoshida, K.; Sakata, Y.; Yoshida, A. Safety and Effectiveness of Lenvatinib in 594 Patients with Unresectable Thyroid Cancer in an All-Case Post-Marketing Observational Study in Japan. Adv. Ther. 2020, 37, 3850–3862. [Google Scholar] [CrossRef] [PubMed]
  6. Xu, B.; Fuchs, T.; Dogan, S.; Landa, I.; Katabi, N.; Fagin, J.A.; Tuttle, R.M.; Sherman, E.; Gill, A.J.; Ghossein, R. Dissecting Anaplastic Thyroid Carcinoma: A Comprehensive Clinical, Histologic, Immunophenotypic, and Molecular Study of 360 Cases. Thyroid 2020, 30, 1505–1517. [Google Scholar] [CrossRef] [PubMed]
  7. Bible, K.C.; Kebebew, E.; Brierley, J.; Brito, J.P.; Cabanillas, M.E.; Clark, T.J., Jr.; Di Cristofano, A.; Foote, R.; Giordano, T.; Kasperbauer, J.; et al. 2021 American Thyroid Association Guidelines for Management of Patients with Anaplastic Thyroid Cancer. Thyroid 2021, 31, 337–386. [Google Scholar] [CrossRef] [PubMed]
  8. Filetti, S.; Durante, C.; Hartl, D.; Leboulleux, S.; Locati, L.D.; Newbold, K.; Papotti, M.G.; Berruti, A. Thyroid cancer: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up†. Ann. Oncol. 2019, 30, 1856–1883. [Google Scholar] [CrossRef] [PubMed]
  9. Haddad, R.I.; Bischoff, L.; Ball, D.; Bernet, V.; Blomain, E.; Busaidy, N.L.; Campbell, M.; Dickson, P.; Duh, Q.Y.; Ehya, H.; et al. Thyroid Carcinoma, Version 2.2022, NCCN Clinical Practice Guidelines in Oncology. J. Natl. Compr. Cancer Netw. 2022, 20, 925–951. [Google Scholar] [CrossRef]
  10. Subbiah, V.; Kreitman, R.J.; Wainberg, Z.A.; Cho, J.Y.; Schellens, J.H.M.; Soria, J.C.; Wen, P.Y.; Zielinski, C.C.; Cabanillas, M.E.; Boran, A.; et al. Dabrafenib plus trametinib in patients with BRAF V600E-mutant anaplastic thyroid cancer: Updated analysis from the phase II ROAR basket study. Ann. Oncol. 2022, 33, 406–415. [Google Scholar] [CrossRef]
  11. De Leo, S.; Trevisan, M.; Fugazzola, L. Recent advances in the management of anaplastic thyroid cancer. Thyroid Res. 2020, 13, 17. [Google Scholar] [CrossRef]
  12. Jannin, A.; Escande, A.; Al Ghuzlan, A.; Blanchard, P.; Hartl, D.; Chevalier, B.; Deschamps, F.; Lamartina, L.; Lacroix, L.; Dupuy, C.; et al. Anaplastic Thyroid Carcinoma: An Update. Cancers 2022, 14, 1061. [Google Scholar] [CrossRef]
  13. Abe, I.; Lam, A.K. Anaplastic Thyroid Carcinoma: Current Issues in Genomics and Therapeutics. Curr. Oncol. Rep. 2021, 23, 31. [Google Scholar] [CrossRef] [PubMed]
  14. Chiacchio, S.; Lorenzoni, A.; Boni, G.; Rubello, D.; Elisei, R.; Mariani, G. Anaplastic thyroid cancer: Prevalence, diagnosis and treatment. Minerva Endocrinol. 2008, 33, 341–357. [Google Scholar] [PubMed]
  15. Califano, I.; Smulever, A.; Jerkovich, F.; Pitoia, F. Advances in the management of anaplastic thyroid carcinoma: Transforming a life-threatening condition into a potentially treatable disease. Rev. Endocr. Metab. Disord. 2024, 25, 123–147. [Google Scholar] [CrossRef]
  16. Shaha, A.R. Implications of prognostic factors and risk groups in the management of differentiated thyroid cancer. Laryngoscope 2004, 114, 393–402. [Google Scholar] [CrossRef] [PubMed]
  17. Maniakas, A.; Dadu, R.; Busaidy, N.L.; Wang, J.R.; Ferrarotto, R.; Lu, C.; Williams, M.D.; Gunn, G.B.; Hofmann, M.C.; Cote, G.; et al. Evaluation of Overall Survival in Patients With Anaplastic Thyroid Carcinoma, 2000–2019. JAMA Oncol. 2020, 6, 1397–1404. [Google Scholar] [CrossRef]
  18. Song, T.; Chen, L.; Zhang, H.; Lu, Y.; Yu, K.; Zhan, W.; Fang, M. Multimodal treatment based on thyroidectomy improves survival in patients with metastatic anaplastic thyroid carcinoma: A SEER analysis from 1998 to 2015. Gland Surg. 2020, 9, 1205–1213. [Google Scholar] [CrossRef]
  19. de Ridder, M.; Nieveen van Dijkum, E.; Engelsman, A.; Kapiteijn, E.; Klümpen, H.J.; Rasch, C.R.N. Anaplastic thyroid carcinoma: A nationwide cohort study on incidence, treatment and survival in the Netherlands over 3 decades. Eur. J. Endocrinol. 2020, 183, 203–209. [Google Scholar] [CrossRef]
  20. Tuttle, R.M.; Haugen, B.; Perrier, N.D. Updated American Joint Committee on Cancer/Tumor-Node-Metastasis Staging System for Differentiated and Anaplastic Thyroid Cancer (Eighth Edition): What Changed and Why? Thyroid 2017, 27, 751–756. [Google Scholar] [CrossRef]
  21. Molinaro, E.; Romei, C.; Biagini, A.; Sabini, E.; Agate, L.; Mazzeo, S.; Materazzi, G.; Sellari-Franceschini, S.; Ribechini, A.; Torregrossa, L.; et al. Anaplastic thyroid carcinoma: From clinicopathology to genetics and advanced therapies. Nat. Rev. Endocrinol. 2017, 13, 644–660. [Google Scholar] [CrossRef]
  22. Onoda, N.; Sugitani, I.; Ito, K.I.; Suzuki, A.; Higashiyama, T.; Fukumori, T.; Suganuma, N.; Masudo, K.; Nakayama, H.; Uno, A.; et al. Evaluation of the 8th Edition TNM Classification for Anaplastic Thyroid Carcinoma. Cancers 2020, 12, 552. [Google Scholar] [CrossRef]
  23. Sugitani, I.; Miyauchi, A.; Sugino, K.; Okamoto, T.; Yoshida, A.; Suzuki, S. Prognostic factors and treatment outcomes for anaplastic thyroid carcinoma: ATC Research Consortium of Japan cohort study of 677 patients. World J. Surg. 2012, 36, 1247–1254. [Google Scholar] [CrossRef] [PubMed]
  24. Haymart, M.R.; Banerjee, M.; Yin, H.; Worden, F.; Griggs, J.J. Marginal treatment benefit in anaplastic thyroid cancer. Cancer 2013, 119, 3133–3139. [Google Scholar] [CrossRef] [PubMed]
  25. Akaishi, J.; Sugino, K.; Kitagawa, W.; Nagahama, M.; Kameyama, K.; Shimizu, K.; Ito, K.; Ito, K. Prognostic factors and treatment outcomes of 100 cases of anaplastic thyroid carcinoma. Thyroid 2011, 21, 1183–1189. [Google Scholar] [CrossRef]
  26. Yoo, S.K.; Song, Y.S.; Lee, E.K.; Hwang, J.; Kim, H.H.; Jung, G.; Kim, Y.A.; Kim, S.J.; Cho, S.W.; Won, J.K.; et al. Integrative analysis of genomic and transcriptomic characteristics associated with progression of aggressive thyroid cancer. Nat. Commun. 2019, 10, 2764. [Google Scholar] [CrossRef] [PubMed]
  27. Ravi, N.; Yang, M.; Gretarsson, S.; Jansson, C.; Mylona, N.; Sydow, S.R.; Woodward, E.L.; Ekblad, L.; Wennerberg, J.; Paulsson, K. Identification of Targetable Lesions in Anaplastic Thyroid Cancer by Genome Profiling. Cancers 2019, 11, 402. [Google Scholar] [CrossRef]
  28. Pozdeyev, N.; Gay, L.M.; Sokol, E.S.; Hartmaier, R.; Deaver, K.E.; Davis, S.; French, J.D.; Borre, P.V.; LaBarbera, D.V.; Tan, A.C.; et al. Genetic Analysis of 779 Advanced Differentiated and Anaplastic Thyroid Cancers. Clin. Cancer Res. 2018, 24, 3059–3068. [Google Scholar] [CrossRef]
  29. Khan, S.A.; Ci, B.; Xie, Y.; Gerber, D.E.; Beg, M.S.; Sherman, S.I.; Cabanillas, M.E.; Busaidy, N.L.; Burtness, B.A.; Heilmann, A.M.; et al. Unique mutation patterns in anaplastic thyroid cancer identified by comprehensive genomic profiling. Head Neck 2019, 41, 1928–1934. [Google Scholar] [CrossRef]
  30. Landa, I.; Ibrahimpasic, T.; Boucai, L.; Sinha, R.; Knauf, J.A.; Shah, R.H.; Dogan, S.; Ricarte-Filho, J.C.; Krishnamoorthy, G.P.; Xu, B.; et al. Genomic and transcriptomic hallmarks of poorly differentiated and anaplastic thyroid cancers. J. Clin. Investig. 2016, 126, 1052–1066. [Google Scholar] [CrossRef]
  31. Bonhomme, B.; Godbert, Y.; Perot, G.; Al Ghuzlan, A.; Bardet, S.; Belleannée, G.; Crinière, L.; Do Cao, C.; Fouilloux, G.; Guyetant, S.; et al. Molecular Pathology of Anaplastic Thyroid Carcinomas: A Retrospective Study of 144 Cases. Thyroid 2017, 27, 682–692. [Google Scholar] [CrossRef] [PubMed]
  32. Duan, H.; Li, Y.; Hu, P.; Gao, J.; Ying, J.; Xu, W.; Zhao, D.; Wang, Z.; Ye, J.; Lizaso, A.; et al. Mutational profiling of poorly differentiated and anaplastic thyroid carcinoma by the use of targeted next-generation sequencing. Histopathology 2019, 75, 890–899. [Google Scholar] [CrossRef] [PubMed]
  33. Romei, C.; Tacito, A.; Molinaro, E.; Piaggi, P.; Cappagli, V.; Pieruzzi, L.; Matrone, A.; Viola, D.; Agate, L.; Torregrossa, L.; et al. Clinical, pathological and genetic features of anaplastic and poorly differentiated thyroid cancer: A single institute experience. Oncol. Lett. 2018, 15, 9174–9182. [Google Scholar] [CrossRef] [PubMed]
  34. Marei, H.E.; Althani, A.; Afifi, N.; Hasan, A.; Caceci, T.; Pozzoli, G.; Morrione, A.; Giordano, A.; Cenciarelli, C. p53 signaling in cancer progression and therapy. Cancer Cell Int. 2021, 21, 703. [Google Scholar] [CrossRef] [PubMed]
  35. McKelvey, B.A.; Umbricht, C.B.; Zeiger, M.A. Telomerase Reverse Transcriptase (TERT) Regulation in Thyroid Cancer: A Review. Front. Endocrinol. 2020, 11, 485. [Google Scholar] [CrossRef] [PubMed]
  36. Panebianco, F.; Nikitski, A.V.; Nikiforova, M.N.; Nikiforov, Y.E. Spectrum of TERT promoter mutations and mechanisms of activation in thyroid cancer. Cancer Med. 2019, 8, 5831–5839. [Google Scholar] [CrossRef] [PubMed]
  37. Matsuse, M.; Mitsutake, N. TERT promoter mutations in thyroid cancer. Endocr. J. 2023, 70, 1035–1049. [Google Scholar] [CrossRef] [PubMed]
  38. Akıncılar, S.C.; Chua, J.Y.H.; Ng, Q.F.; Chan, C.H.T.; Eslami, S.Z.; Chen, K.; Low, J.L.; Arumugam, S.; Aswad, L.; Chua, C.; et al. Identification of mechanism of cancer-cell-specific reactivation of hTERT offers therapeutic opportunities for blocking telomerase specifically in human colorectal cancer. Nucleic Acids Res. 2023, 51, 1–16. [Google Scholar] [CrossRef] [PubMed]
  39. Low, K.C.; Tergaonkar, V. Telomerase: Central regulator of all of the hallmarks of cancer. Trends Biochem. Sci. 2013, 38, 426–434. [Google Scholar] [CrossRef]
  40. Chen, S.; Yang, L.; Dong, H.; Guo, H. Human telomerase reverse transcriptase recruits the β-catenin/TCF-4 complex to transactivate chemokine (C-C motif) ligand 2 expression in colorectal cancer. Biomed. Pharmacother. 2019, 112, 108700. [Google Scholar] [CrossRef]
  41. Kumar, N.; Sethi, G. Telomerase and hallmarks of cancer: An intricate interplay governing cancer cell evolution. Cancer Lett. 2023, 578, 216459. [Google Scholar] [CrossRef] [PubMed]
  42. Lavoie, H.; Gagnon, J.; Therrien, M. ERK signalling: A master regulator of cell behaviour, life and fate. Nat. Rev. Mol. Cell Biol. 2020, 21, 607–632. [Google Scholar] [CrossRef] [PubMed]
  43. Luo, W.; Sun, R.; Chen, X.; Li, J.; Jiang, J.; He, Y.; Shi, S.; Wen, H. ERK Activation-Mediated Autophagy Induction Resists Licochalcone A-Induced Anticancer Activities in Lung Cancer Cells in vitro. OncoTargets Ther. 2020, 13, 13437–13450. [Google Scholar] [CrossRef] [PubMed]
  44. Samson, S.C.; Khan, A.M.; Mendoza, M.C. ERK signaling for cell migration and invasion. Front. Mol. Biosci. 2022, 9, 998475. [Google Scholar] [CrossRef] [PubMed]
  45. Schubert, L.; Mariko, M.L.; Clerc, J.; Huillard, O.; Groussin, L. MAPK Pathway Inhibitors in Thyroid Cancer: Preclinical and Clinical Data. Cancers 2023, 15, 710. [Google Scholar] [CrossRef] [PubMed]
  46. Cook, F.A.; Cook, S.J. Inhibition of RAF dimers: It takes two to tango. Biochem. Soc. Trans. 2021, 49, 237–251. [Google Scholar] [CrossRef] [PubMed]
  47. Rosove, M.H.; Peddi, P.F.; Glaspy, J.A. BRAF V600E inhibition in anaplastic thyroid cancer. N. Engl. J. Med. 2013, 368, 684–685. [Google Scholar] [CrossRef] [PubMed]
  48. Rushton, S.; Burghel, G.; Wallace, A.; Nonaka, D. Immunohistochemical detection of BRAF V600E mutation status in anaplastic thyroid carcinoma. Histopathology 2016, 69, 524–526. [Google Scholar] [CrossRef]
  49. Perri, F.; Pezzullo, L.; Chiofalo, M.G.; Lastoria, S.; Di Gennaro, F.; Scarpati, G.D.; Caponigro, F. Targeted therapy: A new hope for thyroid carcinomas. Crit. Rev. Oncol./Hematol. 2015, 94, 55–63. [Google Scholar] [CrossRef]
  50. Hlozek, J.; Pekova, B.; Rotnágl, J.; Holý, R.; Astl, J. Genetic Changes in Thyroid Cancers and the Importance of Their Preoperative Detection in Relation to the General Treatment and Determination of the Extent of Surgical Intervention-A Review. Biomedicines 2022, 10, 1515. [Google Scholar] [CrossRef]
  51. Yamazaki, H.; Kunisaki, C.; Sugimori, M.; Rino, Y.; Saito, A. Genetic landscape of 482 thyroid carcinomas: Analysis with the national datacenter for cancer genomic medicine in Japan. Endocrine 2024, in press. [Google Scholar] [CrossRef] [PubMed]
  52. Rusinek, D.; Chmielik, E.; Krajewska, J.; Jarzab, M.; Oczko-Wojciechowska, M.; Czarniecka, A.; Jarzab, B. Current Advances in Thyroid Cancer Management. Are We Ready for the Epidemic Rise of Diagnoses? Int. J. Mol. Sci. 2017, 18, 1817. [Google Scholar] [CrossRef] [PubMed]
  53. Oishi, N.; Kondo, T.; Ebina, A.; Sato, Y.; Akaishi, J.; Hino, R.; Yamamoto, N.; Mochizuki, K.; Nakazawa, T.; Yokomichi, H.; et al. Molecular alterations of coexisting thyroid papillary carcinoma and anaplastic carcinoma: Identification of TERT mutation as an independent risk factor for transformation. Mod. Pathol. 2017, 30, 1527–1537. [Google Scholar] [CrossRef] [PubMed]
  54. Ferrari, S.M.; Elia, G.; Ragusa, F.; Ruffilli, I.; La Motta, C.; Paparo, S.R.; Patrizio, A.; Vita, R.; Benvenga, S.; Materazzi, G.; et al. Novel treatments for anaplastic thyroid carcinoma. Gland. Surg. 2020, 9, S28–S42. [Google Scholar] [CrossRef] [PubMed]
  55. Gilani, S.M.; Abi-Raad, R.; Garritano, J.; Cai, G.; Prasad, M.L.; Adeniran, A.J. RAS mutation and associated risk of malignancy in the thyroid gland: An FNA study with cytology-histology correlation. Cancer Cytopathol. 2022, 130, 284–293. [Google Scholar] [CrossRef] [PubMed]
  56. Lai, W.A.; Liu, C.Y.; Lin, S.Y.; Chen, C.C.; Hang, J.F. Characterization of Driver Mutations in Anaplastic Thyroid Carcinoma Identifies RAS and PIK3CA Mutations as Negative Survival Predictors. Cancers 2020, 12, 1973. [Google Scholar] [CrossRef] [PubMed]
  57. Haigis, K.M.; Kendall, K.R.; Wang, Y.; Cheung, A.; Haigis, M.C.; Glickman, J.N.; Niwa-Kawakita, M.; Sweet-Cordero, A.; Sebolt-Leopold, J.; Shannon, K.M.; et al. Differential effects of oncogenic K-Ras and N-Ras on proliferation, differentiation and tumor progression in the colon. Nat. Genet. 2008, 40, 600–608. [Google Scholar] [CrossRef] [PubMed]
  58. Kunstman, J.W.; Juhlin, C.C.; Goh, G.; Brown, T.C.; Stenman, A.; Healy, J.M.; Rubinstein, J.C.; Choi, M.; Kiss, N.; Nelson-Williams, C.; et al. Characterization of the mutational landscape of anaplastic thyroid cancer via whole-exome sequencing. Hum. Mol. Genet. 2015, 24, 2318–2329. [Google Scholar] [CrossRef] [PubMed]
  59. Krishnamoorthy, G.P.; Davidson, N.R.; Leach, S.D.; Zhao, Z.; Lowe, S.W.; Lee, G.; Landa, I.; Nagarajah, J.; Saqcena, M.; Singh, K.; et al. EIF1AX and RAS Mutations Cooperate to Drive Thyroid Tumorigenesis through ATF4 and c-MYC. Cancer Discov. 2019, 9, 264–281. [Google Scholar] [CrossRef]
  60. Janku, F.; Yap, T.A.; Meric-Bernstam, F. Targeting the PI3K pathway in cancer: Are we making headway? Nat. Rev. Clin. Oncol. 2018, 15, 273–291. [Google Scholar] [CrossRef]
  61. Glaviano, A.; Foo, A.S.C.; Lam, H.Y.; Yap, K.C.H.; Jacot, W.; Jones, R.H.; Eng, H.; Nair, M.G.; Makvandi, P.; Geoerger, B.; et al. PI3K/AKT/mTOR signaling transduction pathway and targeted therapies in cancer. Mol. Cancer 2023, 22, 138. [Google Scholar] [CrossRef] [PubMed]
  62. Yang, J.; Nie, J.; Ma, X.; Wei, Y.; Peng, Y.; Wei, X. Targeting PI3K in cancer: Mechanisms and advances in clinical trials. Mol. Cancer 2019, 18, 26. [Google Scholar] [CrossRef] [PubMed]
  63. Martini, M.; De Santis, M.C.; Braccini, L.; Gulluni, F.; Hirsch, E. PI3K/AKT signaling pathway and cancer: An updated review. Ann. Med. 2014, 46, 372–383. [Google Scholar] [CrossRef] [PubMed]
  64. Mayer, I.A.; Arteaga, C.L. The PI3K/AKT Pathway as a Target for Cancer Treatment. Annu. Rev. Med. 2016, 67, 11–28. [Google Scholar] [CrossRef] [PubMed]
  65. Deng, J.; Bai, X.; Feng, X.; Ni, J.; Beretov, J.; Graham, P.; Li, Y. Inhibition of PI3K/Akt/mTOR signaling pathway alleviates ovarian cancer chemoresistance through reversing epithelial-mesenchymal transition and decreasing cancer stem cell marker expression. BMC Cancer 2019, 19, 618. [Google Scholar] [CrossRef] [PubMed]
  66. Peng, Y.; Wang, Y.; Zhou, C.; Mei, W.; Zeng, C. PI3K/Akt/mTOR Pathway and Its Role in Cancer Therapeutics: Are We Making Headway? Front. Oncol. 2022, 12, 819128. [Google Scholar] [CrossRef]
  67. Samuels, Y.; Diaz, L.A., Jr.; Schmidt-Kittler, O.; Cummins, J.M.; Delong, L.; Cheong, I.; Rago, C.; Huso, D.L.; Lengauer, C.; Kinzler, K.W.; et al. Mutant PIK3CA promotes cell growth and invasion of human cancer cells. Cancer Cell 2005, 7, 561–573. [Google Scholar] [CrossRef]
  68. Thorpe, L.M.; Yuzugullu, H.; Zhao, J.J. PI3K in cancer: Divergent roles of isoforms, modes of activation and therapeutic targeting. Nat. Rev. Cancer 2015, 15, 7–24. [Google Scholar] [CrossRef]
  69. Pinto, N.; Ruicci, K.M.; Khan, M.I.; Shaikh, M.H.; Zeng, Y.F.P.; Yoo, J.; Fung, K.; MacNeil, S.D.; Mendez, A.; Mymryk, J.S.; et al. Introduction and expression of PIK3CA(E545K) in a papillary thyroid cancer BRAF(V600E) cell line leads to a dedifferentiated aggressive phenotype. J. Otolaryngol. Head Neck Surg. 2022, 51, 7. [Google Scholar] [CrossRef]
  70. Worby, C.A.; Dixon, J.E. PTEN. Annu. Rev. Biochem. 2014, 83, 641–669. [Google Scholar] [CrossRef]
  71. Hashemi, M.; Taheriazam, A.; Daneii, P.; Hassanpour, A.; Kakavand, A.; Rezaei, S.; Hejazi, E.S.; Aboutalebi, M.; Gholamrezaie, H.; Saebfar, H.; et al. Targeting PI3K/Akt signaling in prostate cancer therapy. J. Cell Commun. Signal. 2023, 17, 423–443. [Google Scholar] [CrossRef] [PubMed]
  72. Paskeh, M.D.A.; Ghadyani, F.; Hashemi, M.; Abbaspour, A.; Zabolian, A.; Javanshir, S.; Razzazan, M.; Mirzaei, S.; Entezari, M.; Goharrizi, M.; et al. Biological impact and therapeutic perspective of targeting PI3K/Akt signaling in hepatocellular carcinoma: Promises and Challenges. Pharmacol. Res. 2023, 187, 106553. [Google Scholar] [CrossRef] [PubMed]
  73. Puzio-Kuter, A.M.; Castillo-Martin, M.; Kinkade, C.W.; Wang, X.; Shen, T.H.; Matos, T.; Shen, M.M.; Cordon-Cardo, C.; Abate-Shen, C. Inactivation of p53 and Pten promotes invasive bladder cancer. Genes Dev. 2009, 23, 675–680. [Google Scholar] [CrossRef] [PubMed]
  74. Jin, Z.; Cheng, X.; Feng, H.; Kuang, J.; Yang, W.; Peng, C.; Shen, B.; Qiu, W. Apatinib Inhibits Angiogenesis Via Suppressing Akt/GSK3β/ANG Signaling Pathway in Anaplastic Thyroid Cancer. Cell. Physiol. Biochem. 2017, 44, 1471–1484. [Google Scholar] [CrossRef] [PubMed]
  75. Prete, A.; Matrone, A.; Gambale, C.; Torregrossa, L.; Minaldi, E.; Romei, C.; Ciampi, R.; Molinaro, E.; Elisei, R. Poorly Differentiated and Anaplastic Thyroid Cancer: Insights into Genomics, Microenvironment and New Drugs. Cancers 2021, 13, 3200. [Google Scholar] [CrossRef] [PubMed]
  76. Toda, S.; Hiroshima, Y.; Iwasaki, H.; Masudo, K. Genomic landscape and clinical features of advanced thyroid carcinoma: A national database study in Japan. J. Clin. Endocrinol. Metab. 2024, dgae271. [Google Scholar] [CrossRef] [PubMed]
  77. Lee, H.; Kim, S.Y.; Kim, S.M.; Chang, H.J.; Lee, Y.S.; Park, C.S.; Chang, H.S. Long-term survival of patients with anaplastic thyroid cancer after multimodal treatment. Transl. Cancer Res. 2020, 9, 5430–5436. [Google Scholar] [CrossRef]
  78. Guo, H.; Lin, H. The value of multimodal treatment in anaplastic thyroid cancer patients with distant metastasis. BMC Surg. 2024, 24, 79. [Google Scholar] [CrossRef] [PubMed]
  79. Hvilsom, G.B.; Londero, S.C.; Hahn, C.H.; Schytte, S.; Pedersen, H.B.; Christiansen, P.; Kiss, K.; Larsen, S.R.; Jespersen, M.L.; Lelkaitis, G.; et al. Anaplastic thyroid carcinoma in Denmark 1996-2012: A national prospective study of 219 patients. Cancer Epidemiol. 2018, 53, 65–71. [Google Scholar] [CrossRef]
  80. Fan, D.; Ma, J.; Bell, A.C.; Groen, A.H.; Olsen, K.S.; Lok, B.H.; Leeman, J.E.; Anderson, E.; Riaz, N.; McBride, S.; et al. Outcomes of multimodal therapy in a large series of patients with anaplastic thyroid cancer. Cancer 2020, 126, 444–452. [Google Scholar] [CrossRef]
  81. Prasongsook, N.; Kumar, A.; Chintakuntlawar, A.V.; Foote, R.L.; Kasperbauer, J.; Molina, J.; Garces, Y.; Ma, D.; Wittich, M.A.N.; Rubin, J.; et al. Survival in Response to Multimodal Therapy in Anaplastic Thyroid Cancer. J. Clin. Endocrinol. Metab. 2017, 102, 4506–4514. [Google Scholar] [CrossRef] [PubMed]
  82. Iyer, P.C.; Dadu, R.; Ferrarotto, R.; Busaidy, N.L.; Habra, M.A.; Zafereo, M.; Gross, N.; Hess, K.R.; Gule-Monroe, M.; Williams, M.D.; et al. Real-World Experience with Targeted Therapy for the Treatment of Anaplastic Thyroid Carcinoma. Thyroid 2018, 28, 79–87. [Google Scholar] [CrossRef] [PubMed]
  83. Hanna, G.J.; Busaidy, N.L.; Chau, N.G.; Wirth, L.J.; Barletta, J.A.; Calles, A.; Haddad, R.I.; Kraft, S.; Cabanillas, M.E.; Rabinowits, G.; et al. Genomic Correlates of Response to Everolimus in Aggressive Radioiodine-refractory Thyroid Cancer: A Phase II Study. Clin. Cancer Res. 2018, 24, 1546–1553. [Google Scholar] [CrossRef]
  84. Takahashi, S.; Kiyota, N.; Yamazaki, T.; Chayahara, N.; Nakano, K.; Inagaki, L.; Toda, K.; Enokida, T.; Minami, H.; Imamura, Y.; et al. A Phase II study of the safety and efficacy of lenvatinib in patients with advanced thyroid cancer. Future Oncol. 2019, 15, 717–726. [Google Scholar] [CrossRef] [PubMed]
  85. Capdevila, J.; Wirth, L.J.; Ernst, T.; Ponce Aix, S.; Lin, C.C.; Ramlau, R.; Butler, M.O.; Delord, J.P.; Gelderblom, H.; Ascierto, P.A.; et al. PD-1 Blockade in Anaplastic Thyroid Carcinoma. J. Clin. Oncol. 2020, 38, 2620–2627. [Google Scholar] [CrossRef]
  86. Moreno, F.; Reyes, C.; Pineda, C.A.; Castellanos, G.; Cálix, F.; Calderón, J.; Vasquez-Bonilla, W.O. Anaplastic thyroid carcinoma with unusual long-term survival: A case report. J. Med. Case Rep. 2022, 16, 39. [Google Scholar] [CrossRef]
  87. Sugitani, I.; Hasegawa, Y.; Sugasawa, M.; Tori, M.; Higashiyama, T.; Miyazaki, M.; Hosoi, H.; Orita, Y.; Kitano, H. Super-radical surgery for anaplastic thyroid carcinoma: A large cohort study using the Anaplastic Thyroid Carcinoma Research Consortium of Japan database. Head Neck 2014, 36, 328–333. [Google Scholar] [CrossRef]
  88. Hermanek, P.; Wittekind, C. Residual tumor (R) classification and prognosis. Semin. Surg. Oncol. 1994, 10, 12–20. [Google Scholar] [CrossRef]
  89. Wittekind, C.; Compton, C.; Quirke, P.; Nagtegaal, I.; Merkel, S.; Hermanek, P.; Sobin, L.H. A uniform residual tumor (R) classification: Integration of the R classification and the circumferential margin status. Cancer 2009, 115, 3483–3488. [Google Scholar] [CrossRef]
  90. Pierie, J.P.; Muzikansky, A.; Gaz, R.D.; Faquin, W.C.; Ott, M.J. The effect of surgery and radiotherapy on outcome of anaplastic thyroid carcinoma. Ann. Surg. Oncol. 2002, 9, 57–64. [Google Scholar] [CrossRef]
  91. Chang, H.S.; Nam, K.H.; Chung, W.Y.; Park, C.S. Anaplastic thyroid carcinoma: A therapeutic dilemma. Yonsei Med. J. 2005, 46, 759–764. [Google Scholar] [CrossRef] [PubMed]
  92. Oliinyk, D.; Augustin, T.; Rauch, J.; Koehler, V.F.; Belka, C.; Spitzweg, C.; Käsmann, L. Role of surgery to the primary tumor in metastatic anaplastic thyroid carcinoma: Pooled analysis and SEER-based study. J. Cancer Res. Clin. Oncol. 2023, 149, 3527–3547. [Google Scholar] [CrossRef] [PubMed]
  93. Huang, N.S.; Shi, X.; Lei, B.W.; Wei, W.J.; Lu, Z.W.; Yu, P.C.; Wang, Y.; Ji, Q.H.; Wang, Y.L. An Update of the Appropriate Treatment Strategies in Anaplastic Thyroid Cancer: A Population-Based Study of 735 Patients. Int. J. Endocrinol. 2019, 2019, 8428547. [Google Scholar] [CrossRef] [PubMed]
  94. Jacob, J.; Vordermark, D.; Lorenz, K.; Medenwald, D. Prognostic factors in radiotherapy of anaplastic thyroid carcinoma: A single center study over 31 years. Radiat. Oncol. 2023, 18, 71. [Google Scholar] [CrossRef] [PubMed]
  95. Saeed, N.A.; Kelly, J.R.; Deshpande, H.A.; Bhatia, A.K.; Burtness, B.A.; Judson, B.L.; Mehra, S.; Edwards, H.A.; Yarbrough, W.G.; Peter, P.R.; et al. Adjuvant external beam radiotherapy for surgically resected, nonmetastatic anaplastic thyroid cancer. Head Neck 2020, 42, 1031–1044. [Google Scholar] [CrossRef]
  96. Nachalon, Y.; Stern-Shavit, S.; Bachar, G.; Shvero, J.; Limon, D.; Popovtzer, A. Aggressive Palliation and Survival in Anaplastic Thyroid Carcinoma. JAMA Otolaryngol. Head Neck Surg. 2015, 141, 1128–1132. [Google Scholar] [CrossRef] [PubMed]
  97. Pezzi, T.A.; Mohamed, A.S.R.; Sheu, T.; Blanchard, P.; Sandulache, V.C.; Lai, S.Y.; Cabanillas, M.E.; Williams, M.D.; Pezzi, C.M.; Lu, C.; et al. Radiation therapy dose is associated with improved survival for unresected anaplastic thyroid carcinoma: Outcomes from the National Cancer Data Base. Cancer 2017, 123, 1653–1661. [Google Scholar] [CrossRef] [PubMed]
  98. Tennvall, J.; Lundell, G.; Wahlberg, P.; Bergenfelz, A.; Grimelius, L.; Akerman, M.; Hjelm Skog, A.L.; Wallin, G. Anaplastic thyroid carcinoma: Three protocols combining doxorubicin, hyperfractionated radiotherapy and surgery. Br. J. Cancer 2002, 86, 1848–1853. [Google Scholar] [CrossRef]
  99. De Crevoisier, R.; Baudin, E.; Bachelot, A.; Leboulleux, S.; Travagli, J.P.; Caillou, B.; Schlumberger, M. Combined treatment of anaplastic thyroid carcinoma with surgery, chemotherapy, and hyperfractionated accelerated external radiotherapy. Int. J. Radiat. Oncol. Biol. Phys. 2004, 60, 1137–1143. [Google Scholar] [CrossRef]
  100. Wang, Y.; Tsang, R.; Asa, S.; Dickson, B.; Arenovich, T.; Brierley, J. Clinical outcome of anaplastic thyroid carcinoma treated with radiotherapy of once- and twice-daily fractionation regimens. Cancer 2006, 107, 1786–1792. [Google Scholar] [CrossRef]
  101. Dandekar, P.; Harmer, C.; Barbachano, Y.; Rhys-Evans, P.; Harrington, K.; Nutting, C.; Newbold, K. Hyperfractionated Accelerated Radiotherapy (HART) for anaplastic thyroid carcinoma: Toxicity and survival analysis. Int. J. Radiat. Oncol. Biol. Phys. 2009, 74, 518–521. [Google Scholar] [CrossRef]
  102. Oliinyk, D.; Augustin, T.; Koehler, V.F.; Rauch, J.; Belka, C.; Spitzweg, C.; Käsmann, L. Hypofractionated Radiotherapy for Anaplastic Thyroid Cancer: Systematic Review and Pooled Analysis. Cancers 2020, 12, 2506. [Google Scholar] [CrossRef] [PubMed]
  103. Stavas, M.J.; Shinohara, E.T.; Attia, A.; Ning, M.S.; Friedman, J.M.; Cmelak, A.J. Short course high dose radiotherapy in the treatment of anaplastic thyroid carcinoma. J. Thyroid Res. 2014, 2014, 764281. [Google Scholar] [CrossRef]
  104. Takahashi, N.; Matsushita, H.; Umezawa, R.; Yamamoto, T.; Ishikawa, Y.; Katagiri, Y.; Tasaka, S.; Takeda, K.; Fukui, K.; Kadoya, N.; et al. Hypofractionated Radiotherapy for Anaplastic Thyroid Carcinoma: 15 Years of Experience in a Single Institution. Eur. Thyroid J. 2019, 8, 24–30. [Google Scholar] [CrossRef]
  105. Nutting, C.M.; Convery, D.J.; Cosgrove, V.P.; Rowbottom, C.; Vini, L.; Harmer, C.; Dearnaley, D.P.; Webb, S. Improvements in target coverage and reduced spinal cord irradiation using intensity-modulated radiotherapy (IMRT) in patients with carcinoma of the thyroid gland. Radiother. Oncol. 2001, 60, 173–180. [Google Scholar] [CrossRef]
  106. Verbakel, W.F.; Cuijpers, J.P.; Hoffmans, D.; Bieker, M.; Slotman, B.J.; Senan, S. Volumetric intensity-modulated arc therapy vs. conventional IMRT in head-and-neck cancer: A comparative planning and dosimetric study. Int. J. Radiat. Oncol. Biol. Phys. 2009, 74, 252–259. [Google Scholar] [CrossRef]
  107. Sherman, E.J.; Harris, J.; Bible, K.C.; Xia, P.; Ghossein, R.A.; Chung, C.H.; Riaz, N.; Gunn, G.B.; Foote, R.L.; Yom, S.S.; et al. Radiotherapy and paclitaxel plus pazopanib or placebo in anaplastic thyroid cancer (NRG/RTOG 0912): A randomised, double-blind, placebo-controlled, multicentre, phase 2 trial. Lancet Oncol. 2023, 24, 175–186. [Google Scholar] [CrossRef] [PubMed]
  108. Lee, N.Y.; Riaz, N.; Wu, V.; Brinkman, T.; Tsai, C.J.; Zhi, W.; Fetten, J.; Ho, A.; Wong, R.J.; Ghossein, R.; et al. A Pilot Study of Durvalumab (MEDI4736) with Tremelimumab in Combination with Image-Guided Stereotactic Body Radiotherapy in the Treatment of Metastatic Anaplastic Thyroid Cancer. Thyroid 2022, 32, 799–806. [Google Scholar] [CrossRef] [PubMed]
  109. Paganetti, H.; Niemierko, A.; Ancukiewicz, M.; Gerweck, L.E.; Goitein, M.; Loeffler, J.S.; Suit, H.D. Relative biological effectiveness (RBE) values for proton beam therapy. Int. J. Radiat. Oncol. Biol. Phys. 2002, 53, 407–421. [Google Scholar] [CrossRef]
  110. Matsumoto, Y.; Fukumitsu, N.; Ishikawa, H.; Nakai, K.; Sakurai, H. A Critical Review of Radiation Therapy: From Particle Beam Therapy (Proton, Carbon, and BNCT) to Beyond. J. Pers. Med. 2021, 11, 825. [Google Scholar] [CrossRef]
  111. Sekihara, K.; Himuro, H.; Saito, N.; Ota, Y.; Kouro, T.; Kusano, Y.; Minohara, S.; Hirayama, R.; Katoh, H.; Sasada, T.; et al. Evaluation of X-ray and carbon-ion beam irradiation with chemotherapy for the treatment of cervical adenocarcinoma cells in 2D and 3D cultures. Cancer Cell Int. 2022, 22, 391. [Google Scholar] [CrossRef]
  112. Youssef, I.; Yoon, J.; Mohamed, N.; Zakeri, K.; Press, R.H.; Yu, Y.; Kang, J.J.; Wong, R.J.; Tuttle, R.M.; Shaha, A.; et al. Outcomes and Toxicities of Nonmedullary Thyroid Tumors Treated with Proton Beam Radiation Therapy. Int. J. Part. Ther. 2022, 9, 20–30. [Google Scholar] [CrossRef]
  113. Mohamad, O.; Makishima, H.; Kamada, T. Evolution of Carbon Ion Radiotherapy at the National Institute of Radiological Sciences in Japan. Cancers 2018, 10, 66. [Google Scholar] [CrossRef] [PubMed]
  114. Mattke, M.; Vogt, K.; Bougatf, N.; Welzel, T.; Oelmann-Avendano, J.; Hauswald, H.; Jensen, A.; Ellerbrock, M.; Jäkel, O.; Haberer, T.; et al. High control rates of proton- and carbon-ion-beam treatment with intensity-modulated active raster scanning in 101 patients with skull base chondrosarcoma at the Heidelberg Ion Beam Therapy Center. Cancer 2018, 124, 2036–2044. [Google Scholar] [CrossRef]
  115. Huang, Q.; Hu, W.; Hu, J.; Gao, J.; Yang, J.; Qiu, X.; Kong, L.; Lu, J.J. Intensity-modulated proton and carbon-ion radiation therapy in the management of major salivary gland carcinomas. Ann. Transl. Med. 2022, 10, 1195. [Google Scholar] [CrossRef] [PubMed]
  116. Ebner, D.K.; Frank, S.J.; Inaniwa, T.; Yamada, S.; Shirai, T. The Emerging Potential of Multi-Ion Radiotherapy. Front. Oncol. 2021, 11, 624786. [Google Scholar] [CrossRef]
  117. Hirayama, R.; Uzawa, A.; Obara, M.; Takase, N.; Koda, K.; Ozaki, M.; Noguchi, M.; Matsumoto, Y.; Li, H.; Yamashita, K.; et al. Determination of the relative biological effectiveness and oxygen enhancement ratio for micronuclei formation using high-LET radiation in solid tumor cells: An in vitro and in vivo study. Mutat. Res./Genet. Toxicol. Environ. Mutagen. 2015, 793, 41–47. [Google Scholar] [CrossRef]
  118. Ito, A.; Kitabatake, S.; Furuichi, W.; Takase, N.; Nakahara, T.; Akiyama, T.; Yoshida, S.; Kusano, Y.; Furusawa, Y.; Hirayama, R. LET Dependence of 8-Hydroxy-2’-deoxyguanosine (8-OHdG) Generation in Mammalian Cells under Air-Saturated and Hypoxic Conditions: A Possible Experimental Approach to the Mechanism of the Decreasing Oxygen Effect in the High-LET Region. Radiat. Res. 2024, 201, 189–196. [Google Scholar] [CrossRef] [PubMed]
  119. Imaizumi, A.; Hirayama, R.; Ikoma, Y.; Nitta, N.; Obata, T.; Hasegawa, S. Neon ion (20Ne10+) charged particle beams manipulate rapid tumor reoxygenation in syngeneic mouse models. Cancer Sci. 2024, 115, 227–236. [Google Scholar] [CrossRef]
  120. Higashiyama, T.; Ito, Y.; Hirokawa, M.; Fukushima, M.; Uruno, T.; Miya, A.; Matsuzuka, F.; Miyauchi, A. Induction chemotherapy with weekly paclitaxel administration for anaplastic thyroid carcinoma. Thyroid 2010, 20, 7–14. [Google Scholar] [CrossRef]
  121. Troch, M.; Koperek, O.; Scheuba, C.; Dieckmann, K.; Hoffmann, M.; Niederle, B.; Raderer, M. High efficacy of concomitant treatment of undifferentiated (anaplastic) thyroid cancer with radiation and docetaxel. J. Clin. Endocrinol. Metab. 2010, 95, E54–E57. [Google Scholar] [CrossRef] [PubMed]
  122. Hanrahan, A.J.; Chen, Z.; Rosen, N.; Solit, D.B. BRAF—A tumour-agnostic drug target with lineage-specific dependencies. Nat. Rev. Clin. Oncol. 2024, 21, 224–247. [Google Scholar] [CrossRef] [PubMed]
  123. Poulikakos, P.I.; Persaud, Y.; Janakiraman, M.; Kong, X.; Ng, C.; Moriceau, G.; Shi, H.; Atefi, M.; Titz, B.; Gabay, M.T.; et al. RAF inhibitor resistance is mediated by dimerization of aberrantly spliced BRAF(V600E). Nature 2011, 480, 387–390. [Google Scholar] [CrossRef] [PubMed]
  124. Lito, P.; Pratilas, C.A.; Joseph, E.W.; Tadi, M.; Halilovic, E.; Zubrowski, M.; Huang, A.; Wong, W.L.; Callahan, M.K.; Merghoub, T.; et al. Relief of profound feedback inhibition of mitogenic signaling by RAF inhibitors attenuates their activity in BRAFV600E melanomas. Cancer Cell 2012, 22, 668–682. [Google Scholar] [CrossRef] [PubMed]
  125. Hirata, E.; Girotti, M.R.; Viros, A.; Hooper, S.; Spencer-Dene, B.; Matsuda, M.; Larkin, J.; Marais, R.; Sahai, E. Intravital imaging reveals how BRAF inhibition generates drug-tolerant microenvironments with high integrin β1/FAK signaling. Cancer Cell 2015, 27, 574–588. [Google Scholar] [CrossRef] [PubMed]
  126. Kurata, K.; Onoda, N.; Noda, S.; Kashiwagi, S.; Asano, Y.; Hirakawa, K.; Ohira, M. Growth arrest by activated BRAF and MEK inhibition in human anaplastic thyroid cancer cells. Int. J. Oncol. 2016, 49, 2303–2308. [Google Scholar] [CrossRef]
  127. Lorimer, C.; Cheng, L.; Chandler, R.; Garcez, K.; Gill, V.; Graham, K.; Grant, W.; Sardo Infirri, S.; Wadsley, J.; Wall, L.; et al. Dabrafenib and Trametinib Therapy for Advanced Anaplastic Thyroid Cancer—Real-World Outcomes From UK Centres. Clin. Oncol. 2023, 35, e60–e66. [Google Scholar] [CrossRef] [PubMed]
  128. Tahara, M.; Kiyota, N.; Imai, H.; Takahashi, S.; Nishiyama, A.; Tamura, S.; Shimizu, Y.; Kadowaki, S.; Ito, K.I.; Toyoshima, M.; et al. A Phase 2 Study of Encorafenib in Combination with Binimetinib in Patients with Metastatic BRAF-Mutated Thyroid Cancer in Japan. Thyroid 2024, 34, 467–476. [Google Scholar] [CrossRef] [PubMed]
  129. Kim, G.; McKee, A.E.; Ning, Y.M.; Hazarika, M.; Theoret, M.; Johnson, J.R.; Xu, Q.C.; Tang, S.; Sridhara, R.; Jiang, X.; et al. FDA approval summary: Vemurafenib for treatment of unresectable or metastatic melanoma with the BRAFV600E mutation. Clin. Cancer Res. 2014, 20, 4994–5000. [Google Scholar] [CrossRef]
  130. Zhang, L.; Gaskins, K.; Yu, Z.; Xiong, Y.; Merino, M.J.; Kebebew, E. An in vivo mouse model of metastatic human thyroid cancer. Thyroid 2014, 24, 695–704. [Google Scholar] [CrossRef]
  131. Hyman, D.M.; Puzanov, I.; Subbiah, V.; Faris, J.E.; Chau, I.; Blay, J.Y.; Wolf, J.; Raje, N.S.; Diamond, E.L.; Hollebecque, A.; et al. Vemurafenib in Multiple Nonmelanoma Cancers with BRAF V600 Mutations. N. Engl. J. Med. 2015, 373, 726–736. [Google Scholar] [CrossRef] [PubMed]
  132. Li, Q.; Li, Z.; Luo, T.; Shi, H. Targeting the PI3K/AKT/mTOR and RAF/MEK/ERK pathways for cancer therapy. Mol. Biomed. 2022, 3, 47. [Google Scholar] [CrossRef]
  133. Zaballos, M.A.; Acuña-Ruiz, A.; Morante, M.; Riesco-Eizaguirre, G.; Crespo, P.; Santisteban, P. Inhibiting ERK dimerization ameliorates BRAF-driven anaplastic thyroid cancer. Cell. Mol. Life Sci. 2022, 79, 504. [Google Scholar] [CrossRef] [PubMed]
  134. Owonikoko, T.K.; Zhang, G.; Lallani, S.B.; Chen, Z.; Martinson, D.E.; Khuri, F.R.; Lonial, S.; Marcus, A.; Sun, S.Y. Evaluation of preclinical efficacy of everolimus and pasireotide in thyroid cancer cell lines and xenograft models. PLoS ONE 2019, 14, e0206309. [Google Scholar] [CrossRef]
  135. Lim, S.M.; Chang, H.; Yoon, M.J.; Hong, Y.K.; Kim, H.; Chung, W.Y.; Park, C.S.; Nam, K.H.; Kang, S.W.; Kim, M.K.; et al. A multicenter, phase II trial of everolimus in locally advanced or metastatic thyroid cancer of all histologic subtypes. Ann. Oncol. 2013, 24, 3089–3094. [Google Scholar] [CrossRef]
  136. Schneider, T.C.; de Wit, D.; Links, T.P.; van Erp, N.P.; van der Hoeven, J.J.; Gelderblom, H.; Roozen, I.C.; Bos, M.; Corver, W.E.; van Wezel, T.; et al. Everolimus in Patients With Advanced Follicular-Derived Thyroid Cancer: Results of a Phase II Clinical Trial. J. Clin. Endocrinol. Metab. 2017, 102, 698–707. [Google Scholar] [CrossRef] [PubMed]
  137. Shimobayashi, M.; Hall, M.N. Making new contacts: The mTOR network in metabolism and signalling crosstalk. Nat. Rev. Mol. Cell Biol. 2014, 15, 155–162. [Google Scholar] [CrossRef] [PubMed]
  138. Gökmen-Polar, Y.; Liu, Y.; Toroni, R.A.; Sanders, K.L.; Mehta, R.; Badve, S.; Rommel, C.; Sledge, G.W., Jr. Investigational drug MLN0128, a novel TORC1/2 inhibitor, demonstrates potent oral antitumor activity in human breast cancer xenograft models. Breast Cancer Res. Treat. 2012, 136, 673–682. [Google Scholar] [CrossRef]
  139. Harada, M.; Benito, J.; Yamamoto, S.; Kaur, S.; Arslan, D.; Ramirez, S.; Jacamo, R.; Platanias, L.; Matsushita, H.; Fujimura, T.; et al. The novel combination of dual mTOR inhibitor AZD2014 and pan-PIM inhibitor AZD1208 inhibits growth in acute myeloid leukemia via HSF pathway suppression. Oncotarget 2015, 6, 37930–37947. [Google Scholar] [CrossRef]
  140. Sekihara, K.; Saitoh, K.; Han, L.; Ciurea, S.; Yamamoto, S.; Kikkawa, M.; Kazuno, S.; Taka, H.; Kaga, N.; Arai, H.; et al. Targeting mantle cell lymphoma metabolism and survival through simultaneous blockade of mTOR and nuclear transporter exportin-1. Oncotarget 2017, 8, 34552–34564. [Google Scholar] [CrossRef]
  141. Jordan, N.J.; Dutkowski, C.M.; Barrow, D.; Mottram, H.J.; Hutcheson, I.R.; Nicholson, R.I.; Guichard, S.M.; Gee, J.M. Impact of dual mTORC1/2 mTOR kinase inhibitor AZD8055 on acquired endocrine resistance in breast cancer in vitro. Breast Cancer Res. 2014, 16, R12. [Google Scholar] [CrossRef]
  142. Kotian, S.; Zhang, L.; Boufraqech, M.; Gaskins, K.; Gara, S.K.; Quezado, M.; Nilubol, N.; Kebebew, E. Dual Inhibition of HDAC and Tyrosine Kinase Signaling Pathways with CUDC-907 Inhibits Thyroid Cancer Growth and Metastases. Clin. Cancer Res. 2017, 23, 5044–5054. [Google Scholar] [CrossRef]
  143. Hescheler, D.A.; Hartmann, M.J.M.; Riemann, B.; Michel, M.; Bruns, C.J.; Alakus, H.; Chiapponi, C. Anaplastic thyroid cancer: Genome-based search for new targeted therapy options. Endocr. Connect. 2022, 11, e210624. [Google Scholar] [CrossRef] [PubMed]
  144. Massacesi, C.; Di Tomaso, E.; Urban, P.; Germa, C.; Quadt, C.; Trandafir, L.; Aimone, P.; Fretault, N.; Dharan, B.; Tavorath, R.; et al. PI3K inhibitors as new cancer therapeutics: Implications for clinical trial design. OncoTargets Ther. 2016, 9, 203–210. [Google Scholar] [CrossRef]
  145. Lorusso, L.; Newbold, K. Lenvatinib: A new option for the treatment of advanced iodine refractory differentiated thyroid cancer? Future Oncol. 2015, 11, 1719–1727. [Google Scholar] [CrossRef] [PubMed]
  146. Ferrari, S.M.; Bocci, G.; Di Desidero, T.; Elia, G.; Ruffilli, I.; Ragusa, F.; Orlandi, P.; Paparo, S.R.; Patrizio, A.; Piaggi, S.; et al. Lenvatinib exhibits antineoplastic activity in anaplastic thyroid cancer in vitro and in vivo. Oncol. Rep. 2018, 39, 2225–2234. [Google Scholar] [CrossRef]
  147. Koyama, S.; Miyake, N.; Fujiwara, K.; Morisaki, T.; Fukuhara, T.; Kitano, H.; Takeuchi, H. Lenvatinib for Anaplastic Thyroid Cancer and Lenvatinib-Induced Thyroid Dysfunction. Eur. Thyroid J. 2018, 7, 139–144. [Google Scholar] [CrossRef] [PubMed]
  148. Iñiguez-Ariza, N.M.; Ryder, M.M.; Hilger, C.R.; Bible, K.C. Salvage Lenvatinib Therapy in Metastatic Anaplastic Thyroid Cancer. Thyroid 2017, 27, 923–927. [Google Scholar] [CrossRef]
  149. Wirth, L.J.; Brose, M.S.; Sherman, E.J.; Licitra, L.; Schlumberger, M.; Sherman, S.I.; Bible, K.C.; Robinson, B.; Rodien, P.; Godbert, Y.; et al. Open-Label, Single-Arm, Multicenter, Phase II Trial of Lenvatinib for the Treatment of Patients With Anaplastic Thyroid Cancer. J. Clin. Oncol. 2021, 39, 2359–2366. [Google Scholar] [CrossRef]
  150. Higashiyama, T.; Sugino, K.; Hara, H.; Ito, K.I.; Nakashima, N.; Onoda, N.; Tori, M.; Katoh, H.; Kiyota, N.; Ota, I.; et al. Phase II study of the efficacy and safety of lenvatinib for anaplastic thyroid cancer (HOPE). Eur. J. Cancer 2022, 173, 210–218. [Google Scholar] [CrossRef]
  151. Ishihara, S.; Onoda, N.; Noda, S.; Asano, Y.; Tauchi, Y.; Morisaki, T.; Kashiwagi, S.; Takashima, T.; Ohira, M. Sorafenib inhibits vascular endothelial cell proliferation stimulated by anaplastic thyroid cancer cells regardless of BRAF mutation status. Int. J. Oncol. 2019, 55, 1069–1076. [Google Scholar] [CrossRef] [PubMed]
  152. Kim, S.; Yazici, Y.D.; Calzada, G.; Wang, Z.Y.; Younes, M.N.; Jasser, S.A.; El-Naggar, A.K.; Myers, J.N. Sorafenib inhibits the angiogenesis and growth of orthotopic anaplastic thyroid carcinoma xenografts in nude mice. Mol. Cancer Ther. 2007, 6, 1785–1792. [Google Scholar] [CrossRef] [PubMed]
  153. Schiff, B.A.; McMurphy, A.B.; Jasser, S.A.; Younes, M.N.; Doan, D.; Yigitbasi, O.G.; Kim, S.; Zhou, G.; Mandal, M.; Bekele, B.N.; et al. Epidermal growth factor receptor (EGFR) is overexpressed in anaplastic thyroid cancer, and the EGFR inhibitor gefitinib inhibits the growth of anaplastic thyroid cancer. Clin. Cancer Res. 2004, 10, 8594–8602. [Google Scholar] [CrossRef] [PubMed]
  154. Nobuhara, Y.; Onoda, N.; Yamashita, Y.; Yamasaki, M.; Ogisawa, K.; Takashima, T.; Ishikawa, T.; Hirakawa, K. Efficacy of epidermal growth factor receptor-targeted molecular therapy in anaplastic thyroid cancer cell lines. Br. J. Cancer 2005, 92, 1110–1116. [Google Scholar] [CrossRef] [PubMed]
  155. Podtcheko, A.; Ohtsuru, A.; Tsuda, S.; Namba, H.; Saenko, V.; Nakashima, M.; Mitsutake, N.; Kanda, S.; Kurebayashi, J.; Yamashita, S. The selective tyrosine kinase inhibitor, STI571, inhibits growth of anaplastic thyroid cancer cells. J. Clin. Endocrinol. Metab. 2003, 88, 1889–1896. [Google Scholar] [CrossRef] [PubMed]
  156. Di Desidero, T.; Fioravanti, A.; Orlandi, P.; Canu, B.; Giannini, R.; Borrelli, N.; Man, S.; Xu, P.; Fontanini, G.; Basolo, F.; et al. Antiproliferative and proapoptotic activity of sunitinib on endothelial and anaplastic thyroid cancer cells via inhibition of Akt and ERK1/2 phosphorylation and by down-regulation of cyclin-D1. J. Clin. Endocrinol. Metab. 2013, 98, E1465–E1473. [Google Scholar] [CrossRef] [PubMed]
  157. Di Desidero, T.; Orlandi, P.; Gentile, D.; Bocci, G. Effects of Pazopanib Monotherapy vs. Pazopanib and Topotecan Combination on Anaplastic Thyroid Cancer Cells. Front. Oncol. 2019, 9, 1202. [Google Scholar] [CrossRef]
  158. Savvides, P.; Nagaiah, G.; Lavertu, P.; Fu, P.; Wright, J.J.; Chapman, R.; Wasman, J.; Dowlati, A.; Remick, S.C. Phase II trial of sorafenib in patients with advanced anaplastic carcinoma of the thyroid. Thyroid 2013, 23, 600–604. [Google Scholar] [CrossRef] [PubMed]
  159. Ito, Y.; Onoda, N.; Ito, K.I.; Sugitani, I.; Takahashi, S.; Yamaguchi, I.; Kabu, K.; Tsukada, K. Sorafenib in Japanese Patients with Locally Advanced or Metastatic Medullary Thyroid Carcinoma and Anaplastic Thyroid Carcinoma. Thyroid 2017, 27, 1142–1148. [Google Scholar] [CrossRef]
  160. Pennell, N.A.; Daniels, G.H.; Haddad, R.I.; Ross, D.S.; Evans, T.; Wirth, L.J.; Fidias, P.H.; Temel, J.S.; Gurubhagavatula, S.; Heist, R.S.; et al. A phase II study of gefitinib in patients with advanced thyroid cancer. Thyroid 2008, 18, 317–323. [Google Scholar] [CrossRef]
  161. Ha, H.T.; Lee, J.S.; Urba, S.; Koenig, R.J.; Sisson, J.; Giordano, T.; Worden, F.P. A phase II study of imatinib in patients with advanced anaplastic thyroid cancer. Thyroid 2010, 20, 975–980. [Google Scholar] [CrossRef] [PubMed]
  162. Ravaud, A.; de la Fouchardière, C.; Caron, P.; Doussau, A.; Do Cao, C.; Asselineau, J.; Rodien, P.; Pouessel, D.; Nicolli-Sire, P.; Klein, M.; et al. A multicenter phase II study of sunitinib in patients with locally advanced or metastatic differentiated, anaplastic or medullary thyroid carcinomas: Mature data from the THYSU study. Eur. J. Cancer 2017, 76, 110–117. [Google Scholar] [CrossRef] [PubMed]
  163. Bible, K.C.; Suman, V.J.; Menefee, M.E.; Smallridge, R.C.; Molina, J.R.; Maples, W.J.; Karlin, N.J.; Traynor, A.M.; Kumar, P.; Goh, B.C.; et al. A multiinstitutional phase 2 trial of pazopanib monotherapy in advanced anaplastic thyroid cancer. J. Clin. Endocrinol. Metab. 2012, 97, 3179–3184. [Google Scholar] [CrossRef] [PubMed]
  164. Jiang, X.; Wang, J.; Deng, X.; Xiong, F.; Ge, J.; Xiang, B.; Wu, X.; Ma, J.; Zhou, M.; Li, X.; et al. Role of the tumor microenvironment in PD-L1/PD-1-mediated tumor immune escape. Mol. Cancer 2019, 18, 10. [Google Scholar] [CrossRef] [PubMed]
  165. Adam, P.; Kircher, S.; Sbiera, I.; Koehler, V.F.; Berg, E.; Knösel, T.; Sandner, B.; Fenske, W.K.; Bläker, H.; Smaxwil, C.; et al. FGF-Receptors and PD-L1 in Anaplastic and Poorly Differentiated Thyroid Cancer: Evaluation of the Preclinical Rationale. Front. Endocrinol. 2021, 12, 712107. [Google Scholar] [CrossRef] [PubMed]
  166. Chintakuntlawar, A.V.; Rumilla, K.M.; Smith, C.Y.; Jenkins, S.M.; Foote, R.L.; Kasperbauer, J.L.; Morris, J.C.; Ryder, M.; Alsidawi, S.; Hilger, C.; et al. Expression of PD-1 and PD-L1 in Anaplastic Thyroid Cancer Patients Treated With Multimodal Therapy: Results From a Retrospective Study. J. Clin. Endocrinol. Metab. 2017, 102, 1943–1950. [Google Scholar] [CrossRef] [PubMed]
  167. Cantara, S.; Bertelli, E.; Occhini, R.; Regoli, M.; Brilli, L.; Pacini, F.; Castagna, M.G.; Toti, P. Blockade of the programmed death ligand 1 (PD-L1) as potential therapy for anaplastic thyroid cancer. Endocrine 2019, 64, 122–129. [Google Scholar] [CrossRef] [PubMed]
  168. Iyer, P.C.; Dadu, R.; Gule-Monroe, M.; Busaidy, N.L.; Ferrarotto, R.; Habra, M.A.; Zafereo, M.; Williams, M.D.; Gunn, G.B.; Grosu, H.; et al. Salvage pembrolizumab added to kinase inhibitor therapy for the treatment of anaplastic thyroid carcinoma. J. Immunother. Cancer 2018, 6, 68. [Google Scholar] [CrossRef] [PubMed]
  169. Wang, J.R.; Zafereo, M.E.; Dadu, R.; Ferrarotto, R.; Busaidy, N.L.; Lu, C.; Ahmed, S.; Gule-Monroe, M.K.; Williams, M.D.; Sturgis, E.M.; et al. Complete Surgical Resection Following Neoadjuvant Dabrafenib Plus Trametinib in BRAF(V600E)-Mutated Anaplastic Thyroid Carcinoma. Thyroid 2019, 29, 1036–1043. [Google Scholar] [CrossRef]
  170. Boudin, L.; Morvan, J.B.; Thariat, J.; Métivier, D.; Marcy, P.Y.; Delarbre, D. Rationale Efficacy and Safety Evidence of Lenvatinib and Pembrolizumab Association in Anaplastic Thyroid Carcinoma. Curr. Oncol. 2022, 29, 7718–7731. [Google Scholar] [CrossRef]
  171. Kang, Y.K.; Boku, N.; Satoh, T.; Ryu, M.H.; Chao, Y.; Kato, K.; Chung, H.C.; Chen, J.S.; Muro, K.; Kang, W.K.; et al. Nivolumab in patients with advanced gastric or gastro-oesophageal junction cancer refractory to, or intolerant of, at least two previous chemotherapy regimens (ONO-4538-12, ATTRACTION-2): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet 2017, 390, 2461–2471. [Google Scholar] [CrossRef] [PubMed]
  172. Michot, J.M.; Bigenwald, C.; Champiat, S.; Collins, M.; Carbonnel, F.; Postel-Vinay, S.; Berdelou, A.; Varga, A.; Bahleda, R.; Hollebecque, A.; et al. Immune-related adverse events with immune checkpoint blockade: A comprehensive review. Eur. J. Cancer 2016, 54, 139–148. [Google Scholar] [CrossRef] [PubMed]
  173. Patrinely, J.R., Jr.; Johnson, R.; Lawless, A.R.; Bhave, P.; Sawyers, A.; Dimitrova, M.; Yeoh, H.L.; Palmeri, M.; Ye, F.; Fan, R.; et al. Chronic Immune-Related Adverse Events Following Adjuvant Anti-PD-1 Therapy for High-risk Resected Melanoma. JAMA Oncol. 2021, 7, 744–748. [Google Scholar] [CrossRef] [PubMed]
  174. Park, E.M.; Chelvanambi, M.; Bhutiani, N.; Kroemer, G.; Zitvogel, L.; Wargo, J.A. Targeting the gut and tumor microbiota in cancer. Nat. Med. 2022, 28, 690–703. [Google Scholar] [CrossRef] [PubMed]
  175. Jiao, J.; Zheng, Y.; Zhang, Q.; Xia, D.; Zhang, L.; Ma, N. Saliva microbiome changes in thyroid cancer and thyroid nodules patients. Front. Cell. Infect. Microbiol. 2022, 12, 989188. [Google Scholar] [CrossRef] [PubMed]
  176. Kun, Y.; Xiaodong, W.; Haijun, W.; Xiazi, N.; Dai, Q. Exploring the oral-gut microbiota during thyroid cancer: Factors affecting the thyroid functions and cancer development. Food Sci. Nutr. 2023, 11, 5657–5674. [Google Scholar] [CrossRef] [PubMed]
  177. Zhang, J.; Zhang, F.; Zhao, C.; Xu, Q.; Liang, C.; Yang, Y.; Wang, H.; Shang, Y.; Wang, Y.; Mu, X.; et al. Dysbiosis of the gut microbiome is associated with thyroid cancer and thyroid nodules and correlated with clinical index of thyroid function. Endocrine 2019, 64, 564–574. [Google Scholar] [CrossRef] [PubMed]
  178. Liu, C.J.; Chen, S.Q.; Zhang, S.Y.; Wang, J.L.; Tang, X.D.; Yang, K.X.; Li, X.R. The comparison of microbial communities in thyroid tissues from thyroid carcinoma patients. J. Microbiol. 2021, 59, 988–1001. [Google Scholar] [CrossRef]
  179. Dai, D.; Yang, Y.; Yang, Y.; Dang, T.; Xiao, J.; Wang, W.; Teng, L.; Xu, J.; Ye, J.; Jiang, H. Alterations of thyroid microbiota across different thyroid microhabitats in patients with thyroid carcinoma. J. Transl. Med. 2021, 19, 488. [Google Scholar] [CrossRef]
  180. Zheng, L.; Zhang, L.; Tang, L.; Huang, D.; Pan, D.; Guo, W.; He, S.; Huang, Y.; Chen, Y.; Xiao, X.; et al. Gut microbiota is associated with response to (131)I therapy in patients with papillary thyroid carcinoma. Eur. J. Nucl. Med. Mol. Imaging 2023, 50, 1453–1465. [Google Scholar] [CrossRef]
  181. Zhao, L.Y.; Mei, J.X.; Yu, G.; Lei, L.; Zhang, W.H.; Liu, K.; Chen, X.L.; Kołat, D.; Yang, K.; Hu, J.K. Role of the gut microbiota in anticancer therapy: From molecular mechanisms to clinical applications. Signal Transduct. Target. Ther. 2023, 8, 201. [Google Scholar] [CrossRef] [PubMed]
  182. Liu, Q.; Sun, W.; Zhang, H. Interaction of Gut Microbiota with Endocrine Homeostasis and Thyroid Cancer. Cancers 2022, 14, 2656. [Google Scholar] [CrossRef] [PubMed]
  183. Moyer, K.F.; Marcadis, A.R.; Shaha, A.R. Airway management, symptom relief and best supportive care in anaplastic thyroid cancer. Curr. Opin. Otolaryngol. Head Neck Surg. 2020, 28, 74–78. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of the MAPK signaling and PAM signaling pathways in the ATC. Sharp arrows (→) and blunt arrows (┤) indicate promotion and inhibition, respectively.
Figure 1. Schematic illustration of the MAPK signaling and PAM signaling pathways in the ATC. Sharp arrows (→) and blunt arrows (┤) indicate promotion and inhibition, respectively.
Biomedicines 12 01286 g001
Figure 2. Initial treatment of stages IVA, IVB, and IVC ATC. Taken from [7,12], with a few modifications. For each question, blue arrows indicate “Yes,“ and red arrows indicate “No.“ R0: resection for cure or complete remission. R1: complete gross resection; tumor removed grossly but not microscopically.
Figure 2. Initial treatment of stages IVA, IVB, and IVC ATC. Taken from [7,12], with a few modifications. For each question, blue arrows indicate “Yes,“ and red arrows indicate “No.“ R0: resection for cure or complete remission. R1: complete gross resection; tumor removed grossly but not microscopically.
Biomedicines 12 01286 g002
Table 1. Staging and TMN classification of ATC according to the AJCC eighth edition.
Table 1. Staging and TMN classification of ATC according to the AJCC eighth edition.
StageT CategoryN CategoryM Category
IVAT1: Tumor ≤ 2 cm in greatest dimension limited to the thyroid N0:No metastasis to regional nodesM0:No distant metastasis
T2:Tumor > 2 cm but ≤4 cm in greatest dimension limited to the thyroid
T3a:Tumor > 4 cm limited to the thyroid
IVBT1 N1:Metastasis to regional nodesM0
T2
T3a
T3b:Gross extrathyroidal extension invading only strap muscles (sternohyoid, sternothyroid, thyrohyoid, or omohyoid muscles) from a tumor of any sizeN0–N1 M0
T4a:Gross extrathyroidal extension invading subcutaneous soft tissues, larynx, trachea, esophagus, or recurrent laryngeal nerve from a tumor of any size
T4b:Gross extrathyroidal extension invading prevertebral fascia or encasing a carotid artery or mediastinal vessels from a tumor of any size
IVCT1–T4 N0–N1 M1:Distant metastasis
Table 2. Recent studies with multimodal treatment in ATC.
Table 2. Recent studies with multimodal treatment in ATC.
TreatmentStudyNumber of PatientsNumber of PatientsMedian OS (Months)ORR (%)Reference
SurgeryRTCTxMTTIMM
RT (66 Gy)
+ Doxorubicin
(10–20 mg/m2 weekly)
or Paclitaxel
(80 mg/m2 weekly)
Retrospective104
IVA:5
IVB:76
IVC:23
5210199007N.D.Fan et al. (2020) [80]
Doxorubicin + Docetaxel
(20 mg/m2 weekly, each)Carboplatin + Paclitaxel
(50 mg/m2 weekly, each)Doxorubicin only
(20 mg/m2 weekly) Cisplatin only
(30 mg/m2 weekly)
Retrospective30
IVA:2
IVB:22
IVC:6
ND:5
273030002163Prasongsook et al. (2017) [81]
Dabrafenib
(150 mg twice daily)
+ Trametinib (2 mg once daily)
Phase 236
IV:1
IVC:35
30301536414.556Subbiah et al. (2022) [10]
Dabrafenib
(150 mg twice daily)
+ Trametinib (2 mg once daily)
Retrospective16
IVB:4
IVC:12
8791609.350Iyer et al. (2018) [82]
Everolimus (10 mg daily)Phase 27
IVC:7
543704.614Hanna et al. (2018) [83]
Lenvatinib (24 mg daily)Phase 217
IV:17
149717010.624Takahashi et al.(2019) [84]
Spartalizumab
(400 mg every 4 weeks)
Phase 1/242
IV:42
2830254425.919Capdevila et al. (2020) [85]
RT: radiotherapy, CTx: chemotherapy, MTT: molecular targeted therapy, IMM: immunotherapy, OS: overall survival, ORR: overall response rate.
Table 3. Ongoing clinical trials in patients with ATC.
Table 3. Ongoing clinical trials in patients with ATC.
ClinicalTrials.gov IdentifierIntervention/TreatmentPhaseStatus
NCT04552769AbemaciclibPhase 2Active, not recruiting
NCT05453799VudalimabPhase 2Recruiting
NCT04171622Lenvatinib + PembrolizumabPhase 2Recruiting
NCT03975231Dabrafenib + Trametinib + IMRTPhase 1Recruiting
NCT05119296Pembrolizumab (Keytruda)Phase 2Recruiting
NCT04420754AIC100 CAR T CellsPhase 1Recruiting
NCT03449108Aldesleukin (IL2) + Autologous Tumor Infiltrating Lymphocytes LN-145 or LN-145-S1Phase 2Active, not recruiting
NCT03246958Nivolumab + IpilimumabPhase 2Active, not recruiting
NCT04675710Pembrolizumab + Dabrafenib + Trametinib + Surgery + IMRTPhase 2Recruiting
NCT03181100Atezolizumab + Chemotherapy (Cobimetinib, Nab-paclitaxel, Paclitaxel, Vemurafenib)Phase 2Active, not recruiting
NCT03085056Trametinib + PaclitaxelEarly Phase 1Active, not recruiting
NCT04238624Dabrafenib + TrametinibPhase 2Recruiting
NCT04759911Selpercatinib + SurgeryPhase 2Recruiting
NCT06007924Avutometinib + DefactinibPhase 2Recruiting
NCT02041260CabozantinibPhase 2Unknown
NCT04579757Surufatinib + TislelizumabPhase 1/2Active, not recruiting
NCT05059470Pembrolizumab + IMRTPhase 2Recruiting
Table 4. Clinical trials related to radiotherapy in patients with ATC.
Table 4. Clinical trials related to radiotherapy in patients with ATC.
ClinicalTrials.gov IdentifierIntervention/TreatmentPhaseStatusReference
NCT03565536Sorafenib + Surgery + EBRTPhase 2Completed
NCT05659186Tislelizumab +
Anlotinib + RT
Phase 2Recruiting
NCT01236547IMRT + Paclitaxel +
Pazopanib
Phase 2CompletedSherman et al. (2023) [107]
NCT03122496Durvalumab + Tremelimumab +SBRTPhase 1CompletedLee et al. (2022) [108]
NCT03211117Docetaxel + Doxorubicin + IMRT + Pembrolizumab + SurgeryPhase 2Completed
NCT04675710Pembrolizumab +
Dabrafenib + Trametinib + Surgery + IMRT
Phase 2Recruiting
NCT03975231Dabrafenib + Trametinib + IMRTPhase 1Recruiting
NCT00004089Chemotherapy
(Fluorouracil,
Hydroxyurea, Paclitaxel) + Surgery + RT
Phase 2Completed
NCT00077103Chemotherapy (Cisplatin, Doxorubicin) + Fosbretabulin + RTPhase 1/2Terminated
NCT05059470Pembrolizumab + IMRTPhase 2Recruiting
RT: radiotherapy, SBRT: stereotactic body radiotherapy.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sekihara, K.; Himuro, H.; Toda, S.; Saito, N.; Hirayama, R.; Suganuma, N.; Sasada, T.; Hoshino, D. Recent Trends and Potential of Radiotherapy in the Treatment of Anaplastic Thyroid Cancer. Biomedicines 2024, 12, 1286. https://doi.org/10.3390/biomedicines12061286

AMA Style

Sekihara K, Himuro H, Toda S, Saito N, Hirayama R, Suganuma N, Sasada T, Hoshino D. Recent Trends and Potential of Radiotherapy in the Treatment of Anaplastic Thyroid Cancer. Biomedicines. 2024; 12(6):1286. https://doi.org/10.3390/biomedicines12061286

Chicago/Turabian Style

Sekihara, Kazumasa, Hidetomo Himuro, Soji Toda, Nao Saito, Ryoichi Hirayama, Nobuyasu Suganuma, Tetsuro Sasada, and Daisuke Hoshino. 2024. "Recent Trends and Potential of Radiotherapy in the Treatment of Anaplastic Thyroid Cancer" Biomedicines 12, no. 6: 1286. https://doi.org/10.3390/biomedicines12061286

APA Style

Sekihara, K., Himuro, H., Toda, S., Saito, N., Hirayama, R., Suganuma, N., Sasada, T., & Hoshino, D. (2024). Recent Trends and Potential of Radiotherapy in the Treatment of Anaplastic Thyroid Cancer. Biomedicines, 12(6), 1286. https://doi.org/10.3390/biomedicines12061286

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop