Next Article in Journal
Total and Mitochondrial Transcriptomic and Proteomic Insights into Regulation of Bioenergetic Processes for Shoot Fast-Growth Initiation in Moso Bamboo
Next Article in Special Issue
Ectopic Tumor VCAM-1 Expression in Cancer Metastasis and Therapy Resistance
Previous Article in Journal
Distinct Dominant Lineage from In Vitro Expanded Adipose-Derived Stem Cells (ASCs) Exhibits Enhanced Wound Healing Properties
Previous Article in Special Issue
Insights into Modern Therapeutic Approaches in Pediatric Acute Leukemias
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Markers of Pediatric Solid Tumors—Diagnosis, Optimizing Treatments, and Determining Susceptibility: Current State and Future Directions

by
Joanna Trubicka
1,*,
Wiesława Grajkowska
1 and
Bożenna Dembowska-Bagińska
2
1
Department of Pathology, The Children’s Memorial Health Institute, Av. Dzieci Polskich 20, 04-730 Warsaw, Poland
2
Department of Oncology, The Children’s Memorial Health Institute, Av. Dzieci Polskich 20, 04-730 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Cells 2022, 11(7), 1238; https://doi.org/10.3390/cells11071238
Submission received: 27 February 2022 / Revised: 26 March 2022 / Accepted: 29 March 2022 / Published: 6 April 2022
(This article belongs to the Special Issue Harnessing the Immune System to Fight Pediatric Cancer)

Abstract

:
Advances in molecular technologies, from genomics and transcriptomics to epigenetics, are providing unprecedented insight into the molecular landscape of pediatric tumors. Multi-omics approaches provide an opportunity to identify a wide spectrum of molecular alterations that account for the initiation of the neoplastic process in children, response to treatment and disease progression. The detection of molecular markers is crucial to assist clinicians in accurate tumor diagnosis, risk stratification, disease subtyping, prediction of treatment response, and surveillance, allowing also for personalized cancer management. This review summarizes the most recent developments in genomics research and their relevance to the field of pediatric oncology with the aim of generating an overview of the most important, from the clinical perspective, molecular markers for pediatric solid tumors. We present an overview of the molecular markers selected based on therapeutic protocols, guidelines from international committees and scientific societies, and published data.

1. Introduction

Recent decades have witnessed an intensive development of molecular research, which has contributed significantly to a more complete understanding of the molecular basis of childhood cancers. The published studies reveal a high heterogeneity of molecular alterations that account for the initiation of the neoplastic process, response to treatment and disease progression. These findings clearly indicate that the profile of molecular changes occurring in childhood malignancies differs significantly from the one observed in adult cancers. Thus, it is not possible to directly extrapolate the knowledge and experience with molecular markers from adults to the pediatric population. The differences observed pertain not only to the molecular basis, but also to the disease itself, its anatomical site and its histopathological features demanding the development of a different diagnostic and therapeutic approach for this group of patients.
Knowledge of specific childhood cancer genetic alterations present in tumor cells (somatic markers) as well as germline alterations is useful throughout the entire diagnostic and therapeutic process. There are genetic markers specific for histological types of cancer which are necessary for making a correct diagnosis (a group of diagnostic markers). Furthermore, there are molecular markers that correlate to the tumor’s grade, allowing us to predict the course of disease and prognosis (a group of prognostic markers) as well as to predict responses to a specific treatment (a group of predictive markers). Most promising are the molecular markers for targeted therapies. Extensive research on the biology of cancers, including their molecular profile, has influenced the current way of practicing medicine in the field of cancer diagnosis and treatment as well as the development of the so-called “personalized or precision medicine”. In recent years, there has been an attempt to change the paradigm of anticancer treatment, which assumes that the occurrence of specific molecular alterations may determine the efficacy of the administered treatments irrespective of the site and histological type of the tumor.
In addition to the changes occurring in the neoplasm itself, which are used as targets for therapy, alterations identified in the genetic material from patients’ peripheral blood (germline alterations) are also important. They allow us to determine whether the occurrence of a particular cancer is caused by the presence of a molecular defect. If the identified molecular alteration may have been inherited, assessing its presence in other family members enables us to identify individuals at a higher-than-average risk of developing cancer. These individuals should be screened systematically for early cancer detection. Germline alterations can also constitute prognostic and predictive markers.
Childhood cancers are rare, but they remain the second-leading cause of death in the pediatric population. Their incidence is 14–15 cases per 100,000 per year. They account for about 1–1.5% of cancers in the general population. Solid tumors account for approximately 60% of all childhood malignancies. Their broad spectrum includes [1,2,3]:
  • Central nervous system (CNS) tumors (~20–23% *),
  • Neuroblastoma (8–10% *),
  • Wilms tumors, (7–8% *),
  • Malignant bone tumors, (osteosarcoma and Ewing tumor) (~7% *),
  • Soft tissue sarcomas (~7% *),
  • Germ cell tumors (3–6% *),
  • Hepatoblastoma, rarely hepatocarcinoma (0.5–2% *),
  • Retinoblastoma (2.5–3% *),
  • Other malignant epithelial neoplasms and malignant melanomas
  • Other and unspecified carcinomas.
* of all malignant neoplasms of developmental age
Since the mid-1970s, the cure rates for most pediatric solid tumors have increased by as much as 50% [4]. At present, more than 80% of children with cancer are cured. These excellent cure rates are achieved with multidrug chemotherapy combined with surgery and/or radiotherapy in the case of solid tumors. However, there is not much to offer to children with refractory or relapsed disease after first- or second-line treatment. For these patients, innovative and effective medicines are needed.
In recent decades personalized treatments based on molecular markers have been developed for adults with cancer. Nevertheless, not enough progress has been made in the development and authorization of targeted therapies for childhood malignancies. Though molecular markers are routinely used in diagnosis, establishing risk groups in pediatric oncology, and novel medicinal products are being developed, with some exceptions, few breakthrough medicines have come to the market for children. Only few such medicinal products received marketing authorization for the treatment of pediatric malignancies. Among them are ABL-class inhibitors for Philadelphia positive acute lymphoblastic leukemia, anti-CD antibodies for B cell lymphomas, and anti-GD2 antibodies for children with high-risk neuroblastoma. Recently, Larotrectinib for children with NTRK fusion solid tumors and Crizotinib for children and young adults with relapsed or refractory systemic anaplastic large-cell lymphoma received marketing authorization. Phase 2/3 studies with Crizotinib in children and adolescents with recurrent, progressive, and unresectable inflammatory myofibroblastic tumors are forthcoming.
Moreover, genetic data are lacking on the difficulty to treat refractory or relapsed solid tumors, limiting our knowledge of the molecular composition of such entities. To further improve cure rates in pediatric cancer it is essential to continue research and targeted medicine trials using tumor molecular profiling in children and adolescents.
This article describes a spectrum of molecular markers of clinical relevance in pediatric solid tumors. The markers were selected based on therapeutic protocols, guidelines from international committees and scientific societies, and a review of the current literature.

2. Molecular Markers of Clinical Significance in Childhood Solid Tumors

2.1. Central Nervous System Neoplasms

Central nervous system (CNS) neoplasms are the most common solid tumors in children and a leading cause of childhood cancer-related deaths. Moreover, out of all survivors of childhood cancer, patients cured from CNS tumors present with the most severe treatment-related health conditions. At present, CNS neoplasms are the greatest challenge for pediatric oncology in its broad diagnostic and therapeutic aspects.
Nevertheless, it is molecularly the best-understood group of childhood solid tumors. The results of multi-omics research led, amongst other things, to the definition of molecular subgroups in medulloblastoma, which have solid clinical implications (e.g., the WNT-activated medulloblastoma is associated with good prognosis and is the subject of de-escalation therapy trials, reducing late effects of treatment). There are also other CNS tumors with molecularly defined subgroups.
The clinical validity of molecular markers in diagnostic management has already been demonstrated in the WHO 2016 classification of central nervous system tumors, and their role was further emphasized in the guidelines prepared by an international consortium cIMPACT-NOW (The Consortium to Inform Molecular and Practical Approaches to CNS Tumor Taxonomy). In the latest WHO 2021 classification of tumors of the CNS, molecular markers are fundamental for making a proper diagnosis [5].
At the same time, there has been an increase in the number of molecularly targeted drugs in cancer. The selection of optimal therapy absolutely requires the assessment of specific molecular alterations (point mutations, amplifications, fusions, chromosomal rearrangements).
A set of clinically relevant molecular markers for pediatric central nervous system tumors is provided in Table 1 [5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21].

2.2. Neuroblastoma

Neuroblastoma (NBL) is the most frequent pediatric extracranial tumor originating from neural crest progenitor cells. It accounts for approximately 10% of all childhood malignancies and for up to 15% of deaths in children from cancer [22]. It is the most common cancer diagnosed in children under 12 months of age with a median age at diagnosis of 17 months [23]. Neuroblastoma can present along the sympathetic nervous system, with the most common abdominal location in the adrenal gland or sympathetic chain. It is a heterogenous disease which is reflected in its clinical course of spontaneous regression, differentiation or rapid progression despite intensive multimodal treatment. Patients with low- and intermediate-risk neuroblastoma have good prognosis, with cure rates over 85%, while the survival of children with high-risk disease is less than 50%.
The most significant prognostic factors in NBL are the child’s age at diagnosis, disease stage [4], tumor histology, DNA ploidy, MYCN amplification status and the presence of characteristic chromosomal aberrations (Table 2).
Recommendations are based on current therapeutic protocols and diagnostic guidelines from the European Neuroblastoma Group (SIOPEN group) and the INRG (International Neuroblastoma Risk Group Biology Committee) [24].
A set of molecular markers clinically relevant for neuroblastoma is provided in Table 2 [24,25,26,27,28].

2.3. Renal Tumors

Wilms tumor (nephroblastoma, WT) is the most common pediatric renal malignancy, accounting for over 90% of renal tumors. Other less frequently occurring malignancies of the kidney include:
  • clear cell sarcoma of kidney (CCSK),
  • renal cell carcinoma associated with MiTF/TFE translocations,
  • malignant rhabdoid tumor of kidney (MRTK),
  • congenital mesoblastic nephroma (CMN), and others.
The incidence of Wilms tumor is approximately 7 cases per 1 million children younger than 15 years of age, accounting for 5–7 percent of all childhood malignancies. WT can occur in both kidneys (bilateral disease), found in 5–8% of cases. The mean age at diagnosis is 44 months in unilateral cases and 31 months for bilateral cases of Wilms tumor. Wilms tumor is rare in patients older than age 15. A total of 1.5% of cases occur in related family members. Approximately 5% of WTs are associated with known constitutional predisposition syndromes.
Pathogenic changes in suppressors that regulate growth, differentiation and proliferation of embryonic kidney tissue play an essential role in the pathogenesis of this cancer. Alterations in WT1, CTNNB1 or AMER1 (WTX) genes are observed in about one-third of Wilms tumor cases [29,30]. Other important genes that regulate miRNA processing, such as DROSHA, DGCR8, DICER1 and XPO5, are also involved [31,32,33,34]. The third important group consists of certain genes, the expression of which plays a significant role in the early stages of kidney development, such as SIX1 i SIX2, EP300 (CREBBP), MLLT1, BCOR and MYCN. Alterations in the TRIM28 gene are associated with the epithelial subtype of Wilms tumor [35]. In anaplastic Wilms tumors, the presence of TP53 gene alterations is noted. An observed loss of heterozygosity within chromosome regions 1p and 16q as well as an increase in the amount of genetic material within chromosome 1q are associated with an unfavorable prognosis [36,37,38].
Recommendations are based on guidelines from two groups: Children Oncology Group (COG), continuing work of The National Wilms Tumor Study Group (NWTS) and The International Society of Paediatric Oncology-Renal Tumors Study Group (SIOP-RTSG), as well as the UMBRELLA therapeutic protocol and literature data [39,40,41,42].
A set of molecular markers clinically relevant for pediatric renal tumors is listed in Table 3 [35,36,37,43,44,45,46,47].

2.4. Malignant Bone Tumors, including Osteosarcoma and Ewing Sarcoma

Malignant bone tumors account for 6% of all childhood malignancies. The estimated incidence rate is 0.8 per million. The most common are osteosarcoma (56%), followed by Ewing sarcoma (34%) and chondrosarcoma (10%). The diagnosis of primary bone tumors relies on clinicopathological and radiological correlation. The recommended molecular assays include among others assessment of the presence of translocations characteristic of Ewing sarcoma and chondrosarcoma, determination of the status of pathogenic variants in the H3F3A gene that are relevant for the diagnostics of giant cell tumor of soft tissue, as well as assessment of the presence of MDM2, PRIM1, and CDK4 amplifications to differentiate lower grade osteosarcoma. The standard was prepared based on the current literature data [48,49,50,51].
A set of molecular markers clinically relevant for malignant bone tumors of childhood is listed in Table 4 [4,50,51,52,53,54,55,56,57,58,59].

2.5. Soft-Tissue Sarcomas (STS)

Soft-tissue sarcomas are a heterogeneous group of malignant neoplasms that arise from embryonic mesenchymal and neuroectodermal tissue. They include neoplasms of muscle, connective and vascular tissue. Most cases occur in children aged 2–6 years and adolescents older than 12 years. The incidence rate, regardless of gender, ranges from 0.2–1.0/100,000 per year.
The most common STS in the pediatric group (70%) is rhabdomyosarcoma (RMS), with the age of onset usually before 10 years. The most common type of RMS is embryonal RMS. However, it is important to identify the alveolar RMS subtype for therapeutic management due to its worse prognosis. Recently, sclerosing and spindle cell rhabdomyosarcoma was separated as a stand-alone pathologic entity, in which two subtypes are molecularly defined: the infantile subset with VGLL2, TEAD1 and SRF fusion as well as the subset with p.L122R MYOD1 gene pathogenic variant. The presence of MYOD1 alteration is associated with poor outcomes and response to therapy [60].
The remaining neoplasms belong to the non-rhabdomyosarcoma soft tissue sarcoma NRSTS group, which is more common in older children and young adults. Synovial sarcoma and MPNST (malignant peripheral nerve sheath tumor) are the most common neoplasms in this group. Other pathological entities are much less prevalent in children.
The recommendations were prepared based on the current literature data. A set of clinically relevant molecular markers for pediatric soft tissue sarcomas is provided in Table 5 [4,58,59,61].

2.6. Germ Cell Tumors

Germ cell tumors (GCTs) are derived from germ cells at different stages of their differentiation. They present a wide variety of site, histologic, and biological diversity. GCTs account for approximately 3–6% of all malignancies in children. The most common age of onset is between 1 and 6 years. Patients with GCT have good prognosis with cure rates over 85%.
Currently, the recommended molecular assays for germ cell tumors include karyotyping of patients with symptoms of gonadal dysgenesis, gonadoblastoma and mediastinal tumors (for boys) and bilateral ovarian tumors (for girls), in order to detect/rule out the presence of genetic syndromes such as Turner, Swyer, Klinefelter, Fraser, Denys-Drash and others. The literature data also indicate the presence of somatic rearrangements involving chromosomes 1, 6, 11, 12, 16, 20, and 22, the clinical significance of which is currently unknown.
The standards were developed following the guidelines of the French TGM-95 protocol (1995), protocols elaborated by the international expert consortium MaGIC-Malignant Germ Cell International Collaborative and literature data [62].

2.7. Liver Tumors

The most common pediatric liver tumor is hepatoblastoma. The incidence of hepatoblastoma is 2–3 cases per 1 million children under 15 years of age; however, approximately 98% of all cases occur in children under 5 years of age. The mean age at diagnosis is 18 months [63]. Survival rates for children with hepatoblastoma exceed 80%. Molecular alterations reported in hepatoblastoma primarily involve genes, the protein products of which regulate the WNT and NF-κB pathway. Thus, the most common somatic alterations are those in genes: CTNNB1 (80–90%), APC (2–3%), AXIN1, AXIN2 and PIK3CA as well as TERT (2–6%) and NFE2L2 (5–10%) [63,64,65,66,67,68,69]. The results of chromosomal rearrangement analysis indicate that this group of cancers has a higher rate of somatic rearrangements within chromosomes 1, 2, 8, and 20. However, the clinical significance of their occurrence is still unknown [64,65,70]. One of the objectives of the Pediatric Hepatic International Tumor Trial (PHITT (NCT03017326)), which has been running since 2017, is to determine the clinical significance of molecular findings including complete genomic, transcriptomic, and epigenomic profiling for hepatoblastoma patients [71].
Other hepatic tumors of childhood include: hepatocellular carcinoma (HCC) which is very rarely diagnosed in children, usually in older patients (10–14 years), but has been also found in children younger than 5 years of age. The prognosis is dismal, after conservative surgical treatment (30% of children achieving 3-year survival). Liver transplantation in children with HCC contributed to the improvement of overall survival (>70%). The most frequently reported somatic alterations in this cancer are alterations in the TERT and TP53 genes (60% and 25–30%, respectively); however, the molecular background is still unknown [72,73];
  • undifferentiated embryonal sarcoma of the liver (UES) which is a rare liver tumor with onset in children mostly aged between 6 and 10 years. The molecular background of this neoplasm is not fully understood; however, according to literature data, the characteristic features of this neoplasm comprise frequent and extensive chromosome rearrangements, also in the form of chromothripsis [74]. Additionally, alterations were observed within the 19q13.4 region, including a t(11;19) (q13;q13.4) translocation and overexpression of the C19MC region (miRNA cluster). The presence of TP53 gene alterations was also observed [75,76,77,78].
The diagnostic standards for hepatoblastoma according to the International Childhood Liver Tumors Strategy Group (SIOPEL) were included in forming these guidelines. A set of clinically relevant molecular markers for hepatic tumors of childhood age is provided in Table 6 [64,70,78,79,80].

2.8. Retinoblastoma

Retinoblastoma is the most common primary malignant intraocular cancer in children and the second most common cancer of the eye in all age groups after choroidal melanoma. It accounts for 3% of all childhood tumors. The number of cases ranges from 1 in 14,000–1 in 18,000 live births [3,81]. The following forms of retinoblastoma are distinguished as:
  • bilateral or multifocal (25–30% of cases, hereditary form),
  • unilateral or unifocal (70–75% of cases, sporadic form),
  • trilateral form, in which the presence of bilateral disease is accompanied by an embryonic intracranial tumor (pineoblastoma) localized in the midline (4%—only in children with the hereditary form of the disease).
Most cases are diagnosed between the ages of 1 and 3 years. The bilateral form of retinoblastoma is diagnosed earlier, before the age of 1. Rarely, the disease is diagnosed after the age of 5 years.
This cancer is associated with high (85–95%) penetrance RB1 gene alterations [81]. There are recent reports of molecular alterations in retinoblastoma patients in genes other than RB1Table 7 [82,83,84,85].

2.9. Melanoma

Melanoma is a malignant neoplasm of the skin, mucous membranes, or the choroid of the eye originating from melanocytes. The incidence is 1 case per 1 million children under 15 years of age. In the pediatric group, melanoma can present with one of three types:
  • Spitzoid melanoma (SM), the most common form,
  • Melanoma that arises from a congenital melanocytic nevus (CMN)
  • Classic melanoma (“adult-type melanoma”), most similar in terms of causes and risk factors to melanoma diagnosed in adults.
Most commonly, melanoma is associated with molecular changes in genes that regulate the MAPK pathway. The recommendations were prepared following the current literature data. A set of molecular markers clinically relevant for childhood melanoma is provided in Table 8 [86,87,88,89,90,91].

2.10. Ovarian Cancers

The occurrence of ovarian cancer in girls may be associated with syndromes such as DICER1 or RTPS (rhabdoid tumor predisposition syndrome); therefore, both germ cell and somatic alterations are also observed in DICER1 (sertoli-Leydig cell tumors) and SMARCA4 (primary small cell carcinoma of the ovary, hypercalcemic type-SCCOHT) (Table 9) [92].

3. Targeted Treatments for Pediatric Solid Tumours

Treatments for pediatric malignancies have changed vastly over the last several decades and cure rates now reach over 80%. However, there are still children with uncurable malignancies and those who are cured experience treatment related chronic health conditions. The progress in the field of molecular biology, the ability to analyze tissue on genome-wide scales, to identify cancers with specific gene alterations with the intent to develop novel targeted treatments has created new opportunities to further improve survival of childhood cancer patients and their quality of life. To date the use of targeted and immune- therapies in children has been limited. Despite many obstacles of drug development in pediatric oncology some medicinal products have come to the market and are used in front-line treatment. There are ongoing pediatric phase I/II biomarker-driven trials in most difficult to treat solid tumors in children. Table 10 presents selected targeted treatments authorized or in development.

4. Germline Alterations

In recent years, we have been witnessing great progress in understanding the molecular profile of childhood cancers and applying this knowledge to clinical practice. This includes somatic as well as germline alterations. The published findings of different multi-omics studies further highlight the differences between childhood and adult cancers. The global number of somatic alterations, as expressed by the Tumor Mutational Burden (TMB), in childhood malignancies is much lower than in adults. With respect to germline alterations, the opposite is true. The occurrence of childhood cancers is more often determined by the presence of alterations responsible for genetic syndromes. More than 200 such syndromes have been identified, and this number is steadily growing. The fact that congenital cancers are also diagnosed underscores the significance of germline alterations in the pathogenesis of childhood cancers. It is estimated that approximately 7–8% of hematologic malignancies and solid tumors in children are determined by germline alterations [93]. Furthermore, these data appear to be underestimated since germline mosaicism, which is difficult to identify, or epigenetic changes, such as loss of imprinting of the 11p15 region or hypermethylation of the CDKN2A suppressor, are rarely assessed in the routine diagnosis of cancer.
Assessment of the presence of germline alterations is also important in optimizing therapeutic management. If patients present with lesions that result from chromosomal instability or that occur in DNA repair genes, the omittance of radiotherapy or a reduced radiation dose are recommended. One such example is children with choroid plexus carcinoma and germline alterations in TP53 gene. Patients with germline RB1 alterations who underwent radiotherapy have twice the risk of developing secondary cancers compared with patients who did not receive radiotherapy [94]. The presence of germline alterations in the NF1 gene in patients with low-grade gliomas is associated with a better prognosis; hence, the treatment undertaken may be less aggressive [95]. However, patients with germline alterations in genes belonging to the (MMR mismatch repair system) such as MLH1, MSH2, MSH6, and PMS2 with brain cancers require more aggressive chemotherapy [96]. The presence of a germline alterations may also contribute to the earlier cancer onset. Rhabdoid tumors occur more frequently in patients under 4 years of age; however, the mean age of onset in patients with germline SMARCB1 alterations is 6 months [97,98]. Therefore, it seems highly appropriate to introduce the assessment of germinal alteration status into the algorithm of diagnostic and therapeutic management of pediatric cancers.
The following tables (Table 11 and Table 12) present selected genetic syndromes associated with the occurrence of childhood cancers, as well as information about which genes should be assessed in selected cancers of this age group.

5. Material and Conditions for Its Preservation for Genetic Testing

The starting point for most oncogenetic tests used is a tumor tissue sample. Formalin fixation and paraffin embedding (FFPE) is the most common form of tissue preservation; however, for NGS-based protocols, tissue preservation by freezing is much more beneficial. In selected tumors (e.g., neuroblastoma), the impression smear of tumor tissue may serve as the material for genetic assay. To identify germline alterations, the patient’s peripheral blood, saliva and/or buccal swab samples are also collected in addition to the tumor tissue.
Due to the heterogeneous nature of tumors (particularly neuroblastoma and selected gliomas), it is recommended that at least two specimens be collected from the respective tumor tissue. In cases of tumor recurrence and probable changes in the molecular profile of the relapsed tumor tissue, resampling is indicated.
Each tumor tissue specimen for molecular assay should be evaluated for the percentage of tumor cells in the tested specimen. This assessment is routinely performed by a pathologist. If more than one biological sample is available, the most appropriate sample should be selected based on the type of molecular assay planned, the availability of the biological material and the need for it at subsequent stages of diagnostic process.
It is also very important to maintain sterility when collecting material for molecular assays. NGS methods can detect mosaic-type alterations at very low levels. In cases of contamination of the specimen with even a very small amount of material from another patient, false results may be obtained.

6. Future Directions

6.1. Methylation Profile

The rapid development of high-throughput next-generation sequencing methods has significantly contributed to the understanding of the molecular profile of the most common pediatric cancers. Today, we know that, on the one hand, these tumors exhibit a great variety of molecular alterations, but, on the other hand, their total number is small compared to the number of somatic alterations detected in tumors occurring in adults. This is reflected very frequently in the low TMB score, which translates into limited applicability of immunotherapy. Effective immunotherapy can also be limited by the suppressive tumor microenvironment with relatively few effector cells. Generally, this type of therapy in pediatric solid tumors still remains in the early stages of development and significant clinical benefit has yet to be demonstrated.
In contrast, epigenetic changes occupy a special place in the vast spectrum of molecular alterations that are identified in childhood cancers. It seems that this type of alteration is crucial for the initiation of carcinogenesis-related processes. A confirmation of this assumption is the occurrence of different genome DNA methylation patterns in different tumor subtypes. A unique epigenetic signature that represents both the tumor origin and the presence of acquired oncogenic alterations affecting chromatin state constitutes a very promising diagnostic tool for, among others, central nervous system tumors in children. The classification system developed by German National Cancer Institute (DKFZ) in Heidelberg, based on the methylome pattern [99] is slowly becoming a routine tool to accurately classify CNS tumors into distinct molecular subtypes. As a result, it may improve the accuracy of diagnosis and standardize pathomorphological assessment. This diagnostic approach is limited by the requirement for specialized equipment, software, and a large reference database. To address this problem, the DKFZ team developed a free online tool, Classifier [100], which allows for the processing of data obtained in a given laboratory and comparing them to a reference database containing the results of methylation profile analysis from over 2800 cases. An additional advantage of this solution is the ability to verify the histopathological diagnosis in morphologically ambiguous cases and, in the future, also the ability to identify new, very rare tumor subtypes, not only of the central nervous system.

6.2. Liquid Biopsy

There is a growing number of published studies demonstrating the important role of liquid biopsy in the diagnostic and therapeutic management of a number of cancers, including pediatric malignancies. It is a complementary or alternative method to surgical biopsy, as well as a non-invasive, promising tool for early cancer detection, that may also overcome problems of tumor accessibility and heterogeneity of tumor tissue. Various biological fluids, including peripheral blood, urine, cerebrospinal, synovial and ocular space fluids, can be used to obtain such circulating material as tumor cells (CTCs), tumor DNA (cfDNA), RNA (cfRNA), proteins and extracellular vesicles (EVs) for diagnostic assays. Thanks to the advances in technology, it is possible to obtain and analyze such biological material with increased effectiveness. There are data demonstrating the usefulness of liquid biopsy in variety of analysis including small- and large-scale mutation analysis, high throughput sequencing technologies, and analysis of structural or copy number alterations. From a clinical perspective, the results from liquid biopsy can provide reliable data as to the status of the disease and allow us to monitor treatment and to evaluate predictive, prognostic and resistance markers. In some cases, it may be helpful in early detection of recurrence. Thus, liquid biopsy, although still a relatively new method, appears to be a significant application for cancer diagnosis and treatment. The results of published research on neuroblastoma, sarcoma, Wilms tumor, hepatoblastoma and retinoblastoma appear to be very promising [101]. Nevertheless, the implementation of liquid biopsy into clinical practice is still to be completed. Its limitations are mainly due to the lack of standardized, validated methods for such analyses and the rarity and instability of obtained tumor biomolecules. However, researchers agree that liquid biopsy represents a potentially major new method that can be used to detect, monitor and treat cancers. Further studies are required to address the limitations of this technique.

7. Conclusions

The introduction of technologies such as massively parallel DNA sequencing and RNA sequencing, as well as tools for the interpretation of the vast amounts of data obtained with these methods, including bioinformatic or crystallographic methods, creates an opportunity to elucidate the molecular mechanisms of childhood cancers and to develop targeted therapies. Artificial intelligence methods are also becoming increasingly employed to design therapeutic algorithms and identify prognostic and predictive markers [102]. The introduction of monitoring of circulating tumor DNA (ctDNA) using next-generation sequencing will enable future precise monitoring of treatment. The integration of a broad spectrum of data from “-omics” studies provides the basis for the development of cancer-specific classifiers used for precise diagnostics. More novel in vivo and in vitro models as well as 3D cultures are being developed and used to test drugs specifically dedicated to pediatric cancers. The aim of all of these efforts is to identify molecular markers and move them into the clinical setting for more precise diagnosis, risk stratification, and more effective and less toxic treatment in this therapeutically challenging group of patients.

Author Contributions

J.T., writing—original draft preparation, B.D.-B. and W.G.—writing—review and editing, B.D.-B.—supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The study was supported by the National Science Centre, Poland, Grant No 2016/23/B/NZ2/03064 (PI: JT).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare that there are no conflict of interest.

References

  1. Available online: https://seer.cancer.gov/iccc/iccc-iarc-2017.html#fn (accessed on 27 February 2022).
  2. Steliarova-Foucher, E.; Colombet, M.; Ries, L.A.G.; Moreno, F.; Dolya, A.; Bray, F.; Hesseling, P.; Shin, H.Y.; Stiller, C.A.; IICC-3 Contributors. International incidence of childhood cancer, 2001–2010: A population-based registry study. Lancet Oncol. 2017, 18, 719–731. [Google Scholar] [CrossRef]
  3. Styczyński, J.; Balwierz, W.; Dembowska-Bagińska, B.; Kazanowska, B.; Wachowiak, J.; Matysiak, M.; Klukowska, A.; Krawczuk-Rybak, M.; Adamkiewicz-Drożyńska, E.; Młynarski, W.M. Paediatric oncology and haematology in Poland: Position paper. Pediatr. Pol. 2018, 93, 451–461. [Google Scholar] [CrossRef]
  4. Abeloff, M.D.; Armitage, J.O.; Niederhuber, J.E.; Kastan, M.B.; McKenna, W.G. Abeloff’s Clinical Oncology E-Book; Elsevier Health Sciences: Amsterdam, The Netherlands, 2008. [Google Scholar]
  5. WHO Classification of Tumours Editorial Board. World Health Organization Classification of Tumours of the Central Nervous System, 5th ed.; International Agency for Research on Cancer: Lyon, France, 2021.
  6. Ellison, D.W.; Hawkins, C.; Jones, D.T.W.; Onar-Thomas, A.; Pfister, S.M.; Reifenberger, G.; Louis, D.N. cIMPACT-NOW update 4: Diffuse gliomas characterized by MYB, MYBL1, or FGFR1 alterations or BRAFV600E mutation. Acta Neuropathol. 2019, 137, 683–687. [Google Scholar] [CrossRef] [PubMed]
  7. Chiang, J.; Harreld, J.H.; Tinkle, C.L.; Moreira, D.C.; Li, X.; Acharya, S.; Qaddoumi, I.; Ellison, D.W. A single-center study of the clinicopathologic correlates of gliomas with a MYB or MYBL1 alteration. Acta Neuropathol. 2019, 138, 1091–1092. [Google Scholar] [CrossRef]
  8. Ryall, S.; Tabori, U.; Hawkins, C. Pediatric low-grade glioma in the era of molecular diagnostics. Acta Neuropathol. Commun. 2020, 8, 30. [Google Scholar] [CrossRef]
  9. Louis, D.N.; Perry, A.; Wesseling, P.; Brat, D.J.; Cree, I.A.; Figarella-Branger, D.; Hawkins, C.; Ng, H.K.; Pfister, S.M.; Reifenberger, G.; et al. The 2021 WHO Classification of Tumors of the Central Nervous System: A summary. Neuro-Oncology 2021, 23, 1231–1251. [Google Scholar] [CrossRef] [PubMed]
  10. Yoshida, Y.; Nobusawa, S.; Nakata, S.; Nakada, M.; Arakawa, Y.; Mineharu, Y.; Sugita, Y.; Yoshioka, T.; Araki, A.; Sato, Y.; et al. CNS high-grade neuroepithelial tumor with BCOR internal tandem duplication: A comparison with its counterparts in the kidney and soft tissue. Brain Pathol. 2018, 28, 710–720. [Google Scholar] [CrossRef] [PubMed]
  11. Gareton, A.; Tauziède-Espariat, A.; Dangouloff-Ros, V.; Roux, A.; Saffroy, R.; Castel, D.; Kergrohen, T.; Fina, F.; Figarella-Branger, D.; Pagès, M.; et al. The histomolecular criteria established for adult anaplastic pilocytic astrocytoma are not applicable to the pediatric population. Acta Neuropathol. 2020, 139, 287–303. [Google Scholar] [CrossRef] [Green Version]
  12. Louis, D.N.; Giannini, C.; Capper, D.; Paulus, W.; Figarella-Branger, D.; Lopes, M.B.; Batchelor, T.T.; Cairncross, J.G.; van den Bent, M.; Wick, W.; et al. cIMPACT-NOW update 2: Diagnostic clarifications for diffuse midline glioma, H3 K27M-mutant and diffuse astrocytoma/anaplastic astrocytoma, IDH-mutant. Acta Neuropathol. 2018, 135, 639–642. [Google Scholar] [CrossRef] [PubMed]
  13. Louis, D.N.; Wesseling, P.; Aldape, K.; Brat, D.J.; Capper, D.; Cree, I.A.; Eberhart, C.; Figarella-Branger, D.; Fouladi, M.; Fuller, G.N.; et al. cIMPACT-NOW update 6: New entity and diagnostic principle recommendations of the cIMPACT-Utrecht meeting on future CNS tumor classification and grading. Brain Pathol. 2020, 30, 844–856. [Google Scholar] [CrossRef] [PubMed]
  14. Ellison, D.W.; Aldape, K.D.; Capper, D.; Fouladi, M.; Gilbert, M.R.; Gilbertson, R.J.; Hawkins, C.; Merchant, T.E.; Pajtler, K.; Venneti, S.; et al. cIMPACT-NOW update 7: Advancing the molecular classification of ependymal tumors. Brain Pathol. 2020, 30, 863–866. [Google Scholar] [CrossRef]
  15. Jünger, S.T.; Andreiuolo, F.; Mynarek, M.; Dörner, E.; Zur Mühlen, A.; Rutkowski, S.; von Bueren, A.O.; Pietsch, T. Ependymomas in infancy: Underlying genetic alterations, histological features, and clinical outcome. Childs Nerv. Syst. 2020, 36, 2693–2700. [Google Scholar] [CrossRef]
  16. Lucas, C.G.; Villanueva-Meyer, J.E.; Whipple, N.; Oberheim Bush, N.A.; Cooney, T.; Chang, S.; McDermott, M.; Berger, M.; Cham, E.; Sun, P.P.; et al. Myxoid glioneuronal tumor, PDGFRA p.K385-mutant: Clinical, radiologic, and histopathologic features. Brain Pathol. 2020, 30, 479–494. [Google Scholar] [CrossRef]
  17. Johnson, D.R.; Giannini, C.; Jenkins, R.B.; Kim, D.K.; Kaufmann, T.J. Plenty of calcification: Imaging characterization of polymorphous low-grade neuroepithelial tumor of the young. Neuroradiology 2019, 61, 1327–1332. [Google Scholar] [CrossRef] [PubMed]
  18. Guerreiro Stucklin, A.S.; Ryall, S.; Fukuoka, K.; Zapotocky, M.; Lassaletta, A.; Li, C.; Bridge, T.; Kim, B.; Arnoldo, A.; Kowalski, P.E.; et al. Alterations in ALK/ROS1/NTRK/MET drive a group of infantile hemispheric gliomas. Nat. Commun. 2019, 10, 4343. [Google Scholar] [CrossRef] [Green Version]
  19. Mondal, G.; Lee, J.C.; Ravindranathan, A.; Villanueva-Meyer, J.E.; Tran, Q.T.; Allen, S.J.; Barreto, J.; Gupta, R.; Doo, P.; Van Ziffle, J.; et al. Pediatric bithalamic gliomas have a distinct epigenetic signature and frequent EGFR exon 20 insertions resulting in potential sensitivity to targeted kinase inhibition. Acta Neuropathol. 2020, 139, 1071–1088. [Google Scholar] [CrossRef]
  20. Kleinschmidt-DeMasters, B.K.; Mulcahy Levy, J.M. H3 K27M-mutant gliomas in adults vs. children share similar histological features and adverse prognosis. Clin. Neuropathol. 2018, 37, 53–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Komori, T. The molecular framework of pediatric-type diffuse gliomas: Shifting toward the revision of the WHO classification of tumors of the central nervous system. Brain Tumor Pathol. 2021, 38, 1–3. [Google Scholar] [CrossRef]
  22. Park, J.R.; Eggert, A.; Caron, H. Neuroblastoma: Biology, prognosis, and treatment. Hematol. Oncol. Clin. N. Am. 2010, 24, 65–86. [Google Scholar] [CrossRef] [PubMed]
  23. London, W.; Castleberry, R.; Matthay, K.; Look, A.; Seeger, R.; Shimada, H.; Thorner, P.; Brodeur, G.; Maris, J.; Reynolds, C. Evidence for an age cutoff greater than 365 days for neuroblastoma risk group stratification in the Children’s Oncology Group. J. Clin. Oncol. 2005, 23, 6459–6465. [Google Scholar] [CrossRef] [PubMed]
  24. European Low and Intermediate Risk Neuroblastoma Protocol: A SIOPEN Study, Version 3.0; High Risk Neuroblastoma Study 1.5 of SIOPEN: 1 June 2011. Available online: https://clinicaltrials.gov/ct2/show/NCT01704716 (accessed on 27 February 2022).
  25. Ambros, P.F.; Ambros, I.M.; SIOP Europe Neuroblastoma Pathology, B.O.; Bone Marrow Group. Pathology and biology guidelines for resectable and unresectable neuroblastic tumors and bone marrow examination guidelines. Med. Pediatr. Oncol. 2001, 37, 492–504. [Google Scholar] [CrossRef]
  26. Ambros, P.F.; Ambros, I.M.; Brodeur, G.M.; Haber, M.; Khan, J.; Nakagawara, A.; Schleiermacher, G.; Speleman, F.; Spitz, R.; London, W.B.; et al. International consensus for neuroblastoma molecular diagnostics: Report from the International Neuroblastoma Risk Group (INRG) Biology Committee. Br. J. Cancer 2009, 100, 1471–1482. [Google Scholar] [CrossRef] [Green Version]
  27. Trigg, R.M.; Turner, S.D. ALK in Neuroblastoma: Biological and Therapeutic Implications. Cancers 2018, 10, 113. [Google Scholar] [CrossRef] [Green Version]
  28. Bresler, S.C.; Weiser, D.A.; Huwe, P.J.; Park, J.H.; Krytska, K.; Ryles, H.; Laudenslager, M.; Rappaport, E.F.; Wood, A.C.; McGrady, P.W.; et al. ALK mutations confer differential oncogenic activation and sensitivity to ALK inhibition therapy in neuroblastoma. Cancer Cell 2014, 26, 682–694. [Google Scholar] [CrossRef] [Green Version]
  29. Wegert, J.; Wittmann, S.; Leuschner, I.; Geissinger, E.; Graf, N.; Gessler, M. WTX inactivation is a frequent, but late event in Wilms tumors without apparent clinical impact. Genes Chromosomes Cancer 2009, 48, 1102–1111. [Google Scholar] [CrossRef]
  30. Ruteshouser, E.C.; Robinson, S.M.; Huff, V. Wilms tumor genetics: Mutations in WT1, WTX, and CTNNB1 account for only about one-third of tumors. Genes Chromosomes Cancer 2008, 47, 461–470. [Google Scholar] [CrossRef] [Green Version]
  31. Walz, A.L.; Ooms, A.; Gadd, S.; Gerhard, D.S.; Smith, M.A.; Guidry Auvil, J.M.; Meerzaman, D.; Chen, Q.R.; Hsu, C.H.; Yan, C.; et al. Recurrent DGCR8, DROSHA, and SIX homeodomain mutations in favorable histology Wilms tumors. Cancer Cell 2015, 27, 286–297. [Google Scholar] [CrossRef] [Green Version]
  32. Wegert, J.; Ishaque, N.; Vardapour, R.; Geörg, C.; Gu, Z.; Bieg, M.; Ziegler, B.; Bausenwein, S.; Nourkami, N.; Ludwig, N.; et al. Mutations in the SIX1/2 pathway and the DROSHA/DGCR8 miRNA microprocessor complex underlie high-risk blastemal type Wilms tumors. Cancer Cell 2015, 27, 298–311. [Google Scholar] [CrossRef] [Green Version]
  33. Rakheja, D.; Chen, K.S.; Liu, Y.; Shukla, A.A.; Schmid, V.; Chang, T.C.; Khokhar, S.; Wickiser, J.E.; Karandikar, N.J.; Malter, J.S.; et al. Somatic mutations in DROSHA and DICER1 impair microRNA biogenesis through distinct mechanisms in Wilms tumours. Nat. Commun. 2014, 2, 4802. [Google Scholar] [CrossRef]
  34. Torrezan, G.T.; Ferreira, E.N.; Nakahata, A.M.; Barros, B.D.; Castro, M.T.; Correa, B.R.; Krepischi, A.C.; Olivieri, E.H.; Cunha, I.W.; Tabori, U.; et al. Recurrent somatic mutation in DROSHA induces microRNA profile changes in Wilms tumour. Nat. Commun. 2014, 5, 4039. [Google Scholar] [CrossRef] [Green Version]
  35. Ooms, A.H.A.G.; Vujanić, G.M.; D’Hooghe, E.; Collini, P.; L’Herminé-Coulomb, A.; Vokuhl, C.; Graf, N.; Heuvel-Eibrink, M.M.V.D.; de Krijger, R.R. Renal Tumors of Childhood-A Histopathologic Pattern-Based Diagnostic Approach. Cancers 2020, 12, 729. [Google Scholar] [CrossRef] [Green Version]
  36. Klamt, B.; Schulze, M.; Thäte, C.; Mares, J.; Goetz, P.; Kodet, R.; Scheulen, W.; Weirich, A.; Graf, N.; Gessler, M. Allele loss in Wilms tumors of chromosome arms 11q, 16q, and 22q correlate with clinicopathological parameters. Genes Chromosomes Cancer 1998, 22, 287–294. [Google Scholar] [CrossRef]
  37. Grundy, P.E.; Breslow, N.E.; Li, S.; Perlman, E.; Beckwith, J.B.; Ritchey, M.L.; Shamberger, R.C.; Haase, G.M.; D’Angio, G.J.; Donaldson, M.; et al. Loss of heterozygosity for chromosomes 1p and 16q is an adverse prognostic factor in favorable-histology Wilms tumor: A report from the National Wilms Tumor Study Group. J. Clin. Oncol. 2005, 23, 7312–7321. [Google Scholar] [CrossRef]
  38. Bown, N.; Cotterill, S.J.; Roberts, P.; Griffiths, M.; Larkins, S.; Hibbert, S.; Middleton, H.; Kelsey, A.; Tritton, D.; Mitchell, C. Cytogenetic abnormalities and clinical outcome in Wilms tumor: A study by the U.K. cancer cytogenetics group and the U.K. Children’s Cancer Study Group. Med. Pediatr. Oncol. 2002, 38, 11–21. [Google Scholar] [CrossRef] [PubMed]
  39. De Kraker, J.; Graf, N.; Pritchard-Jones, K.; Pein, F. Nephroblastoma Clinical Trial and Study SIOP 2001, Protocol. SIOP RTSG. 2001. Available online: https://www.skion.nl/workspace/uploads/Protocol-SIOP-2001.pdf (accessed on 26 February 2022).
  40. Pietras, W. Advances and changes in the treatment of children with nephroblastoma. Adv. Clin. Exp. Med. 2012, 21, 809–820. [Google Scholar] [PubMed]
  41. Godzinski, J. The current status of treatment of Wilms’ tumor as per the SIOP trials. J. Indian Assoc. Pediatr. Surg. 2015, 20, 16–20. [Google Scholar] [CrossRef] [Green Version]
  42. van den Heuvel-Eibrink, M.M.; Hol, J.A.; Pritchard-Jones, K.; van Tinteren, H.; Furtwängler, R.; Verschuur, A.C.; Vujanic, G.M.; Leuschner, I.; Brok, J.; Rübe, C.; et al. Position paper: Rationale for the treatment of Wilms tumour in the UMBRELLA SIOP-RTSG 2016 protocol. Nat. Rev. Urol. 2017, 14, 743–752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Huff, V.; Miwa, H.; Haber, D.A.; Call, K.M.; Housman, D.; Strong, L.C.; Saunders, G.F. Evidence for WT1 as a Wilms tumor (WT) gene: Intragenic germinal deletion in bilateral WT. Am. J. Hum. Genet 1991, 48, 997–1003. [Google Scholar]
  44. Bardeesy, N.; Falkoff, D.; Petruzzi, M.J.; Nowak, N.; Zabel, B.; Adam, M.; Aguiar, M.C.; Grundy, P.; Shows, T.; Pelletier, J. Anaplastic Wilms’ tumour, a subtype displaying poor prognosis, harbours p53 gene mutations. Nat. Genet 1994, 7, 91–97. [Google Scholar] [CrossRef]
  45. Ooms, A.H.; Gadd, S.; Gerhard, D.S.; Smith, M.A.; Guidry Auvil, J.M.; Meerzaman, D.; Chen, Q.R.; Hsu, C.H.; Yan, C.; Nguyen, C.; et al. Significance of TP53 Mutation in Wilms Tumors with Diffuse Anaplasia: A Report from the Children’s Oncology Group. Clin. Cancer Res. 2016, 22, 5582–5591. [Google Scholar] [CrossRef] [Green Version]
  46. Wegert, J.; Vokuhl, C.; Collord, G.; Del Castillo Velasco-Herrera, M.; Farndon, S.J.; Guzzo, C.; Jorgensen, M.; Anderson, J.; Slater, O.; Duncan, C.; et al. Recurrent intragenic rearrangements of EGFR and BRAF in soft tissue tumors of infants. Nat. Commun. 2018, 9, 2378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Vokuhl, C.; Nourkami-Tutdibi, N.; Furtwängler, R.; Gessler, M.; Graf, N.; Leuschner, I. ETV6-NTRK3 in congenital mesoblastic nephroma: A report of the SIOP/GPOH nephroblastoma study. Pediatr. Blood Cancer 2018, 65, e26925. [Google Scholar] [CrossRef] [PubMed]
  48. Scotlandi, K.; Hattinger, C.M.; Pellegrini, E.; Gambarotti, M.; Serra, M. Genomics and Therapeutic Vulnerabilities of Primary Bone Tumors. Cells 2020, 9, 968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Behjati, S.; Tarpey, P.S.; Presneau, N.; Scheipl, S.; Pillay, N.; Van Loo, P.; Wedge, D.C.; Cooke, S.L.; Gundem, G.; Davies, H.; et al. Distinct H3F3A and H3F3B driver mutations define chondroblastoma and giant cell tumor of bone. Nat. Genet. 2013, 45, 1479–1482. [Google Scholar] [CrossRef] [PubMed]
  50. Mejia-Guerrero, S.; Quejada, M.; Gokgoz, N.; Gill, M.; Parkes, R.K.; Wunder, J.S.; Andrulis, I.L. Characterization of the 12q15 MDM2 and 12q13-14 CDK4 amplicons and clinical correlations in osteosarcoma. Genes Chromosomes Cancer 2010, 49, 518–525. [Google Scholar] [CrossRef] [PubMed]
  51. Casali, P.G.; Bielack, S.; Abecassis, N.; Aro, H.T.; Bauer, S.; Biagini, R.; Bonvalot, S.; Boukovinas, I.; Bovee, J.V.M.G.; Brennan, B.; et al. Bone sarcomas: ESMO-PaedCan-EURACAN Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 2018, 29, iv79–iv95. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, X.; Bahrami, A.; Pappo, A.; Easton, J.; Dalton, J.; Hedlund, E.; Ellison, D.; Shurtleff, S.; Wu, G.; Wei, L.; et al. Recurrent somatic structural variations contribute to tumorigenesis in pediatric osteosarcoma. Cell Rep. 2014, 7, 104–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Ognjanovic, S.; Olivier, M.; Bergemann, T.L.; Hainaut, P. Sarcomas in TP53 germline mutation carriers: A review of the IARC TP53 database. Cancer 2012, 118, 1387–1396. [Google Scholar] [CrossRef] [PubMed]
  54. Heinsohn, S.; Evermann, U.; Zur Stadt, U.; Bielack, S.; Kabisch, H. Determination of the prognostic value of loss of heterozygosity at the retinoblastoma gene in osteosarcoma. Int. J. Oncol. 2007, 30, 1205–1214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Schrage, Y.M.; Lam, S.; Jochemsen, A.G.; Cleton-Jansen, A.M.; Taminiau, A.H.; Hogendoorn, P.C.; Bovée, J.V. Central chondrosarcoma progression is associated with pRb pathway alterations: CDK4 down-regulation and p16 overexpression inhibit cell growth in vitro. J. Cell Mol. Med. 2009, 13, 2843–2852. [Google Scholar] [CrossRef] [PubMed]
  56. El Beaino, M.; Roszik, J.; Livingston, J.A.; Wang, W.L.; Lazar, A.J.; Amini, B.; Subbiah, V.; Lewis, V.; Conley, A.P. Mesenchymal Chondrosarcoma: A Review with Emphasis on its Fusion-Driven Biology. Curr. Oncol. Rep. 2018, 20, 37. [Google Scholar] [CrossRef] [PubMed]
  57. Perry, J.A.; Kiezun, A.; Tonzi, P.; Van Allen, E.M.; Carter, S.L.; Baca, S.C.; Cowley, G.S.; Bhatt, A.S.; Rheinbay, E.; Pedamallu, C.S.; et al. Complementary genomic approaches highlight the PI3K/mTOR pathway as a common vulnerability in osteosarcoma. Proc. Natl. Acad. Sci. USA 2014, 111, E5564–E5573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Fiedorowicz, M.; Bartnik, E.; Sobczuk, P.; Teterycz, P.; Czarnecka, A.M. Molecular biology of sarcoma. Oncol. Clin. Pract. 2018, 14, 307–330. [Google Scholar] [CrossRef]
  59. Schaefer, I.M.; Hong, K.; Kalbasi, A. How Technology Is Improving the Multidisciplinary Care of Sarcoma. Am. Soc. Clin. Oncol. Educ. Book 2020, 40, 445–462. [Google Scholar] [CrossRef] [PubMed]
  60. Agaram, N.P.; LaQuaglia, M.P.; Alaggio, R.; Zhang, L.; Fujisawa, Y.; Ladanyi, M.; Wexler, L.H.; Antonescu, C.R. MYOD1-mutant spindle cell and sclerosing rhabdomyosarcoma: An aggressive subtype irrespective of age. A reappraisal for molecular classification and risk stratification. Mod. Pathol. 2019, 32, 27–36. [Google Scholar] [CrossRef] [PubMed]
  61. Available online: https://www.pathologyoutlines.com/topic/softtissuemolecular.html (accessed on 27 February 2022).
  62. Olson, T.A.; Murray, M.J.; Rodriguez-Galindo, C.; Nicholson, J.C.; Billmire, D.F.; Krailo, M.D.; Dang, H.M.; Amatruda, J.F.; Thornton, C.M.; Arul, G.S.; et al. Pediatric and Adolescent Extracranial Germ Cell Tumors: The Road to Collaboration. J. Clin. Oncol. 2015, 33, 3018–3028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Czauderna, P.; Lopez-Terrada, D.; Hiyama, E.; Häberle, B.; Malogolowkin, M.H.; Meyers, R.L. Hepatoblastoma state of the art: Pathology, genetics, risk stratification, and chemotherapy. Curr. Opin. Pediatr. 2014, 26, 19–28. [Google Scholar] [CrossRef] [PubMed]
  64. Eichenmüller, M.; Trippel, F.; Kreuder, M.; Beck, A.; Schwarzmayr, T.; Häberle, B.; Cairo, S.; Leuschner, I.; von Schweinitz, D.; Strom, T.M.; et al. The genomic landscape of hepatoblastoma and their progenies with HCC-like features. J. Hepatol. 2014, 61, 1312–1320. [Google Scholar] [CrossRef]
  65. Simon, M.M.; Azkargorta, L.G.; Nonell, M.L.; López-Santamaria, M.; Garrido, M.E.; Mateos, C.B.; Plaza, F.D.; Hernandez, M.; Mendiola, I.; Ojanguren, M.; et al. Armengol. Comphrehensive proteomic study of hepatoblastoma: Identification of a prognostic signature and deregulated pathway. J. Hepatol. 2016, 64, S302. [Google Scholar] [CrossRef]
  66. Aguiar, T.F.M.; Rivas, M.P.; Costa, S.; Maschietto, M.; Rodrigues, T.; Sobral de Barros, J.; Barbosa, A.C.; Valieris, R.; Fernandes, G.R.; Bertola, D.R. Insights into the somatic mutation burden of hepatoblastomas from Brazilian patients. Front. Oncol. 2020, 10, 556. [Google Scholar] [CrossRef] [PubMed]
  67. Sumazin, P.; Chen, Y.; Treviño, L.R.; Sarabia, S.F.; Hampton, O.A.; Patel, K.; Mistretta, T.A.; Zorman, B.; Thompson, P.; Heczey, A.; et al. Genomic analysis of hepatoblastoma identifies distinct molecular and prognostic subgroups. Hepatology 2017, 65, 104–121. [Google Scholar] [CrossRef] [Green Version]
  68. Bell, D.; Ranganathan, S.; Tao, J.; Monga, S.P. Novel Advances in Understanding of Molecular Pathogenesis of Hepatoblastoma: A Wnt/β-Catenin Perspective. Gene Expr. 2017, 17, 141–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Czauderna, P.; Haeberle, B.; Hiyama, E.; Rangaswami, A.; Krailo, M.; Maibach, R.; Rinaldi, E.; Feng, Y.; Aronson, D.; Malogolowkin, M.; et al. The Children’s Hepatic tumors International Collaboration (CHIC): Novel global rare tumor database yields new prognostic factors in hepatoblastoma and becomes a research model. Eur. J. Cancer 2016, 52, 92–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Cairo, S.; Armengol, C.; De Reyniès, A.; Wei, Y.; Thomas, E.; Renard, C.A.; Goga, A.; Balakrishnan, A.; Semeraro, M.; Gresh, L.; et al. Hepatic stem-like phenotype and interplay of Wnt/beta-catenin and Myc signaling in aggressive childhood liver cancer. Cancer Cell 2008, 14, 471–484. [Google Scholar] [CrossRef]
  71. Available online: https://clinicaltrials.gov/ct2/history/NCT03017326?V_4=View (accessed on 27 February 2022).
  72. Kelly, D.; Sharif, K.; Brown, R.M.; Morland, B. Hepatocellular Carcinoma in Children. Clin. Liver Dis. 2015, 19, 433–447. [Google Scholar] [CrossRef]
  73. Available online: https://www.orpha.net/consor/cgi-bin/OC_Exp.php?lng=en&Expert=33402 (accessed on 27 February 2022).
  74. Forment, J.V.; Kaidi, A.; Jackson, S.P. Chromothripsis and cancer: Causes and consequences of chromosome shattering. Nat. Rev. Cancer 2012, 12, 663–670. [Google Scholar] [CrossRef] [PubMed]
  75. Lepreux, S.; Rebouissou, S.; Le Bail, B.; Saric, J.; Balabaud, C.; Bloch, B.; Martin-Négrier, M.L.; Zucman-Rossi, J.; Bioulac-Sage, P. Mutation of TP53 gene is involved in carcinogenesis of hepatic undifferentiated (embryonal) sarcoma of the adult, in contrast with Wnt or telomerase pathways: An immunohistochemical study of three cases with genomic relation in two cases. J. Hepatol. 2005, 42, 424–429. [Google Scholar] [CrossRef]
  76. Hu, X.; Chen, H.; Jin, M.; Wang, X.; Lee, J.; Xu, W.; Zhang, R.; Li, S.; Niu, J. Molecular cytogenetic characterization of undifferentiated embryonal sarcoma of the liver: A case report and literature review. Mol. Cytogenet. 2012, 5, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Sangkhathat, S.; Kusafuka, T.; Nara, K.; Yoneda, A.; Fukuzawa, M. Non-random p53 mutations in pediatric undifferentiated (embryonal) sarcoma of the liver. Hepatol. Res. 2006, 35, 229–234. [Google Scholar] [CrossRef] [PubMed]
  78. Zhang, H.; Lei, L.; Zuppan, C.W.; Raza, A.S. Undifferentiated embryonal sarcoma of the liver with an unusual presentation: Case report and review of the literature. J. Gastrointest. Oncol. 2016, 7, S100–S106. [Google Scholar] [CrossRef]
  79. Sekiguchi, M.; Seki, M.; Kawai, T.; Yoshida, K.; Yoshida, M.; Isobe, T.; Hoshino, N.; Shirai, R.; Tanaka, M.; Souzaki, R.; et al. Integrated multiomics analysis of hepatoblastoma unravels its heterogeneity and provides novel druggable targets. NPJ Precis. Oncol. 2020, 4, 20. [Google Scholar] [CrossRef]
  80. Setty, B.A.; Jinesh, G.G.; Arnold, M.; Pettersson, F.; Cheng, C.H.; Cen, L.; Yoder, S.J.; Teer, J.K.; Flores, E.R.; Reed, D.R.; et al. The genomic landscape of undifferentiated embryonal sarcoma of the liver is typified by C19MC structural rearrangement and overexpression combined with TP53 mutation or loss. PLoS Genet. 2020, 16, e1008642. [Google Scholar] [CrossRef]
  81. Rodriguez-Galindo, C.; Buchsbaum, J.C. Chapter 69-Retinoblastoma. In Clinical Radiation Oncology, 3rd ed.; Gunderson, L.L., Tepper, J.E., Eds.; W.B. Saunders: Philadelphia, PA, USA, 2012; pp. 1455–1469. [Google Scholar]
  82. Knudson, A.G. Mutation and cancer: Statistical study of retinoblastoma. Proc. Natl. Acad. Sci. USA 1971, 68, 820–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Zhang, J.; Benavente, C.A.; McEvoy, J.; Flores-Otero, J.; Ding, L.; Chen, X.; Ulyanov, A.; Wu, G.; Wilson, M.; Wang, J.; et al. A novel retinoblastoma therapy from genomic and epigenetic analyses. Nature 2012, 481, 329–334. [Google Scholar] [CrossRef] [PubMed]
  84. Francis, J.H.; Richards, A.L.; Mandelker, D.L.; Berger, M.F.; Walsh, M.F.; Dunkel, I.J.; Donoghue, M.T.A.; Abramson, D.H. Molecular Changes in Retinoblastoma beyond. Cancers 2021, 13, 149. [Google Scholar] [CrossRef]
  85. Zugbi, S.; Ganiewich, D.; Bhattacharyya, A.; Aschero, R.; Ottaviani, D.; Sampor, C.; Cafferata, E.G.; Mena, M.; Sgroi, M.; Winter, U.; et al. Clinical, Genomic, and Pharmacological Study of. Cancers 2020, 12, 2714. [Google Scholar] [CrossRef] [PubMed]
  86. Wiesner, T.; He, J.; Yelensky, R.; Esteve-Puig, R.; Botton, T.; Yeh, I.; Lipson, D.; Otto, G.; Brennan, K.; Murali, R.; et al. Kinase fusions are frequent in Spitz tumours and spitzoid melanomas. Nat. Commun. 2014, 5, 3116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Quan, V.L.; Zhang, B.; Mohan, L.S.; Shi, K.; Isales, M.C.; Panah, E.; Taxter, T.J.; Beaubier, N.; White, K.; Gerami, P. Activating Structural Alterations in MAPK Genes Are Distinct Genetic Drivers in a Unique Subgroup Of Spitzoid Neoplasms. Am. J. Surg. Pathol. 2019, 43, 538–548. [Google Scholar] [CrossRef] [PubMed]
  88. VandenBoom, T.; Quan, V.L.; Zhang, B.; Garfield, E.M.; Kong, B.Y.; Isales, M.C.; Panah, E.; Igartua, C.; Taxter, T.; Beaubier, N.; et al. Genomic Fusions in Pigmented Spindle Cell Nevus of Reed. Am. J. Surg. Pathol. 2018, 42, 1042–1051. [Google Scholar] [CrossRef] [PubMed]
  89. Gerami, P.; Scolyer, R.A.; Xu, X.; Elder, D.E.; Abraham, R.M.; Fullen, D.; Prieto, V.G.; Leboit, P.E.; Barnhill, R.L.; Cooper, C.; et al. Risk assessment for atypical spitzoid melanocytic neoplasms using FISH to identify chromosomal copy number aberrations. Am. J. Surg. Pathol. 2013, 37, 676–684. [Google Scholar] [CrossRef] [PubMed]
  90. Yang, C.; Gru, A.A.; Dehner, L.P. Common and not so Common Melanocytic Lesions in Children and Adolescents. Pediatr. Dev. Pathol. 2018, 21, 252–270. [Google Scholar] [CrossRef] [PubMed]
  91. Gerami, P.; Jewell, S.S.; Morrison, L.E.; Blondin, B.; Schulz, J.; Ruffalo, T.; Matushek, P.; Legator, M.; Jacobson, K.; Dalton, S.R.; et al. Fluorescence in situ hybridization (FISH) as an ancillary diagnostic tool in the diagnosis of melanoma. Am. J. Surg. Pathol. 2009, 33, 1146–1156. [Google Scholar] [CrossRef]
  92. Lala, S.V.; Strubel, N. Ovarian neoplasms of childhood. Pediatr. Radiol. 2019, 49, 1463–1475. [Google Scholar] [CrossRef]
  93. Sweet-Cordero, E.A.; Biegel, J.A. The genomic landscape of pediatric cancers: Implications for diagnosis and treatment. Science 2019, 363, 1170–1175. [Google Scholar] [CrossRef]
  94. Kamihara, J.; Bourdeaut, F.; Foulkes, W.D.; Molenaar, J.J.; Mossé, Y.P.; Nakagawara, A.; Parareda, A.; Scollon, S.R.; Schneider, K.W.; Skalet, A.H.; et al. Retinoblastoma and Neuroblastoma Predisposition and Surveillance. Clin. Cancer Res. 2017, 23, e98–e106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Helfferich, J.; Nijmeijer, R.; Brouwer, O.F.; Boon, M.; Fock, A.; Hoving, E.W.; Meijer, L.; den Dunnen, W.F.; de Bont, E.S. Neurofibromatosis type 1 associated low grade gliomas: A comparison with sporadic low grade gliomas. Crit. Rev. Oncol. Hematol. 2016, 104, 30–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Bouffet, E.; Larouche, V.; Campbell, B.B.; Merico, D.; de Borja, R.; Aronson, M.; Durno, C.; Krueger, J.; Cabric, V.; Ramaswamy, V.; et al. Immune Checkpoint Inhibition for Hypermutant Glioblastoma Multiforme Resulting from Germline Biallelic Mismatch Repair Deficiency. J. Clin. Oncol. 2016, 34, 2206–2211. [Google Scholar] [CrossRef] [Green Version]
  97. Eaton, K.W.; Tooke, L.S.; Wainwright, L.M.; Judkins, A.R.; Biegel, J.A. Spectrum of SMARCB1/INI1 mutations in familial and sporadic rhabdoid tumors. Pediatr. Blood Cancer 2011, 56, 7–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Bourdeaut, F.; Lequin, D.; Brugières, L.; Reynaud, S.; Dufour, C.; Doz, F.; André, N.; Stephan, J.L.; Pérel, Y.; Oberlin, O.; et al. Frequent hSNF5/INI1 germline mutations in patients with rhabdoid tumor. Clin. Cancer Res. 2011, 17, 31–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Capper, D.; Jones, D.T.W.; Sill, M.; Hovestadt, V.; Schrimpf, D.; Sturm, D.; Koelsche, C.; Sahm, F.; Chavez, L.; Reuss, D.E.; et al. DNA methylation-based classification of central nervous system tumours. Nature 2018, 555, 469–474. [Google Scholar] [CrossRef]
  100. Available online: https://www.molecularneuropathology.org/mnp/ (accessed on 27 February 2022).
  101. Weiser, D.A.; West-Szymanski, D.C.; Fraint, E.; Weiner, S.; Rivas, M.A.; Zhao, C.W.T.; He, C.; Applebaum, M.A. Progress toward liquid biopsies in pediatric solid tumors. Cancer Metastasis Rev. 2019, 38, 553–571. [Google Scholar] [CrossRef] [PubMed]
  102. Kumar, Y.; Gupta, S.; Singla, R.; Hu, Y.-C. A Systematic Review of Artificial Intelligence Techniques in Cancer Prediction and Diagnosis. Arch. Comput. Methods Eng. 2021, 1–28. [Google Scholar] [CrossRef] [PubMed]
Table 1. Molecular markers—central nervous system tumors of childhood.
Table 1. Molecular markers—central nervous system tumors of childhood.
Tumor TypeGenes/Molecular
Profiles Characteristically
Altered
Diagnostic MarkerPrognostic, Predictive Marker,
Target for Therapy
Gliomas, glioneuronal tumors, and neuronal tumors
Pediatric-type diffuse
low-grade gliomas
Diffuse astrocytoma,
MYB- or MYBL1-altered
MYB
MYBL1
IDH-wild type (IDH1, IDH2)
H3-wild type
(H3-3A, HIST1H3B, HIST1H3BC)
+Alterations involving MYB and MYBL1 genes:
favorable prognostic factor
Angiocentric gliomaMYB (usually MYB:QKI)+Favorable prognostic factor
Polymorphous low-grade
neuroepithelial tumor of the young
BRAF
FGFR family
+Potential targets for tyrosine kinase inhibitors
(depending on the alteration detected)
Diffuse low-grade glioma,
MAPK pathway-altered
FGFR1
BRAF
+Potential targets for tyrosine kinase inhibitors
(depending on the alteration detected)
Pediatric-type diffuse
high-grade gliomas
Diffuse midline glioma,
H3 K27-altered
H3-3A, HIST1H3B, HIST1H3BC: p.K28M
TP53
ACVR1
PDGFRA
EGFR
EZHIP
+Pathogenic variants in genes encoding histone H3.3 - an unfavorable prognostic factor.
Potential targets for targeted therapy
(depending on the alteration detected)
Diffuse hemispheric glioma, H3 G34-mutantH3-3A: p.G35R/V
TP53
ATRX
MGMT
+MGMT - a favorable prognostic factor associated with increased sensitivity to temozolomide
Diffuse pediatric-type
high-grade glioma,
H3-wildtype and
IDH-wildtype
(subgroups: pedRTK1,
pedRTK2, pedMYCN)
IDH-wild type (IDH1, IDH2)
H3-wild type (H3-3A, HIST1H3B HIST1H3BC)
PDGFRA
MYCN
EGFR
(methylome)
+Potential therapeutic targets
(depending on the alteration detected)
Infant-type hemispheric gliomaNTRK1/2/3
ALK
ROS1
MET
+Potential target for tyrosine kinase inhibitors
(depending on the alteration engraved)
Circumscribed
astrocytic
gliomas
Pilocytic astrocytomaKIAA1549-BRAF
BRAF
NF1
fusions involving
NTRK1 and NTRK2 genes
+Potential targets for tyrosine kinase inhibitors
(depending on the alteration detected)
KIAA1549-BRAF, BRAF alteration- a favorable
prognostic factor
High-grade astrocytoma with piloid featuresIDH1/IDH2 - wild type
EGFR amplification wild type
MAPK signaling pathway primarily: BRAF (mainly fusions),
NF1
ATRX
CDKN2A/B,
(methylome)
+Potential targets for tyrosine kinase inhibitors
(depending on the alteration detected)
Pleomorphic
xanthoastrocytoma
BRAF
CDKN2A/B
+Potential targets for tyrosine kinase inhibitors
(depending on the alteration detected)
Subependymal giant cell
astrocytoma
TSC1
TSC2
+Potential targets for mTOR inhibitors
Astroblastoma,
MN1-altered
MN1
(primarily a fusion with BEND2)
+Favorable prognostic factor
OtherGangliogliomaMAPK signaling pathway:
BRAF
RAS
FGFR1/2
RAF1
NTRK2
NF1
+Potential targets for tyrosine kinase inhibitors
(depending on the alteration detected)
Desmoplastic infantile
ganglioglioma/desmoplastic infantile astrocytoma
MAPK signaling pathway: primarily BRAF alterations+Potential targets for tyrosine kinase inhibitors
(depending on the alteration detected)
Dysembryoplastic
neuroepithelial tumor
FGFR1+Potential targets for tyrosine kinase inhibitors
Rosette-forming
glioneuronal tumor
FGFR1
PIK3CA
NF1
+Potential therapeutic targets
(depending on the alteration detected)
Myxoid glioneuronal tumorPDFGRA+-
Diffuse leptomeningeal
glioneuronal tumor
MAPK signaling pathway,
primarily a fusion KIAA1549-BRAF,
1p structural rearrangements,
(methylome)
+Potential targets for tyrosine kinase inhibitors
(depending on the alteration detected)
Dysplastic cerebellar
gangliocytoma
(Lhermitte-Duclos disease)
PTEN+-
Extraventricular
neurocytoma
IDH-wildtype (IDH1, IDH2)
FGFR (FGFR1-TACC1 fusion)
+-
Ependymal tumors
Supratentorial
ependymoma
Supratentorial ependymoma, ZFTA fusion-positiveZFTA -RELA+Fusion involving the YAP1 gene - a favorable
prognostic factor
Supratentorial ependymoma, YAP1 fusion-positiveYAP1- MAMLD1+Fusion involving the ZFTA gene - an unfavorable prognostic factor
Posterior fossa
ependymoma
Posterior fossa ependymoma, group PFA global reduction of H3 K27me3
(methylome)
+Unfavorable prognostic factor
Posterior fossa ependymoma, group PFBH3 K27me3
(maintaining methylation levels)
(methylome)
+Favorable prognostic factors
Spinal ependymomaSpinal ependymoma,
MYCN-amplified
NF2
MYCN
+MYCN - an unfavorable prognostic factor
Choroid plexus tumors
Choroid plexus carcinomaTP53-Unfavorable prognostic factor associated
with reduced
indications for radiotherapy
Embryonal tumors
Medulloblastomas,
molecularly defined
Medulloblastoma,
WNT-activated
CTNNB1
APC
+Favorable prognostic factors
Medulloblastoma,
SHH-activated and
TP53-wildtype
TP53- wild type
PTCH1
SUFU
SMO
MYCN
GLI2
(methylome)
+Potential targets for SHH pathway inhibitors.
Unfavorable prognostic factors (MYCN)
Medulloblastoma,
SHH-activated and
TP53-mutant
TP53
PTCH1
SUFU
SMO
MYCN
GLI2
(methylome)
+Potential targets for SHH pathway inhibitors.
Unfavorable prognostic factors (TP53, MYCN)
Medulloblastoma,
non-WNT/non-SHH
MYC
MYCN
PRDM6
(methylome)
+
Other CNS embryonal
tumors
Atypical teratoid/rhabdoid
tumor
SMARCB1
SMARCA4
rearrangements of chromosome 22
+SMARCB1 - unfavorable prognostic factor,
SMARCA4 - a standard prognostic factor
Embryonal tumor with
multilayered rosettes
C19MC- DICER1+-
CNS neuroblastoma, FOXR2-activatedFOXR2+-
CNS tumor with BCOR
internal tandem duplication
BCOR+unfavorable prognostic factor
Pineal tumors
PineoblastomaRB1
DICER1
+-
Desmoplastic myxoid
tumor of the pineal
region, SMARCB1-mutant
SMARCB1+-
Table 2. Molecular markers—neuroblastoma.
Table 2. Molecular markers—neuroblastoma.
Tumor TypeGenes/Molecular
Profiles
Characteristically Altered
Diagnostic MarkerPrognostic, Predictive Markers,
Target for Therapy
NeuroblastomaMYCN (amplification)-Unfavorable prognostic factor in patients older than 18 months at diagnosis. The presence of MYCN gene amplification is associated with a significantly higher risk of recurrence and death from progression.
(NCA)—numerical changes in the number of chromosomes in the genetic material of cancer cells -Diploidy as observed in the genetic material of the tumor tissue is associated with an unfavorable course of the disease. In infants, hyperploidy is a favorable prognostic factor (it is associated with good response to chemotherapy).
(SCA)—segmental chromosomal changes most commonly involving chromosome regions 1p, 1q, 2p, 3p, 4p, 11q and 17q -Most frequently observed in advanced stages of the disease in older children, unfavorable prognostic factors.
ALK:
-
SNP(most frequent: p.F1174L, p.F1245C, p.R1275Q)
-
amplification
-
fusions
-A potential target for ALK kinase inhibitors,
unfavorable prognostic factor.
Table 3. Molecular markers—renal tumors of childhood.
Table 3. Molecular markers—renal tumors of childhood.
Tumor TypeGenes/Molecular
Profiles
Characteristically Altered
Diagnostic MarkerPrognostic, Predictive Markers
Wilms tumor (nephroblastoma)
-
WT1 (11p13)
+No clear data
-
LOH 1p/16q
-
TP53
-Unfavorable prognostic factor
Congenital mesoblastic nephroma (CMN)
-
t(12;15)
ETV6-NTRK3
-
EGRF-ITD
+-
Clear cell sarcoma of kidney (CCSK)
-
BCOR- ITDs
-
t(10;17)(q22;p13)
YWHAE-NUTM2B
-
t(12;22)(q13;q12)
EWSR1-AFT1
-
BCOR-CCNB3
+-
Renal carcinoma associated with MiTF/TFE translocations
-
t(X;1)(p11.2;q21.2)
TFE3-PRCC
-
t(X;17)(p11.2;q25)
TFE3-ASPL (ASPSCR1)
-
t(X;1)(p11.2;p34)
TFE3-SFPQ (PSF)
and others
+-
Malignant rhabdoid tumor of kidney (MRTK)
-
SMARCB1
-
SMARCA4
+Unfavorable
prognostic factor
Metanephric tumors
-
BRAF (p.V600E)
+-
Table 4. Molecular markers—malignant bone tumors.
Table 4. Molecular markers—malignant bone tumors.
Tumor TypeGenes/Molecular Profiles Characteristically Altered Diagnostic Marker
OsteosarcomaTP53
RB1
8q21-24 (amplification)
MDM2 (amplification)
extensive and comprehensive
chromosomal rearrangements
+
Ewing sarcomat(11;22)(q24;q12)
EWSR1-FLI1
t(21;22)(q12;q12)
EWSR1-ERG
t(2;22)(q33;q12)
EWSR1-CREB1
t(7;22)(p22;q12)
EWSR1-ETV1
t(17;22)(q12;q12)
EWSR1-E1AF
inv(22)(q12;q12)
EWSR1-ZSG
t(16;21)(p11;q22)
FUS-ERG
and others
+
ChondrosarcomaHEY1-NCOA2
t(1;5)(q42;q32)
RF2BP2-CDX1
IDH1
IDH2
TP53
+
Giant cell tumor of soft tissueH3F3A
HRAS
TP53
+
Table 5. Molecular markers—soft tissue sarcomas.
Table 5. Molecular markers—soft tissue sarcomas.
Tumor TypeGenes/Molecular Profiles Characteristically Altered Diagnostic Marker
Rhabdomyosarcoma
Alveolar
t(2;13)(q35;q14)
PAX3-FOXO1
t(1;13)(p36;q14)
PAX7-FOXO1
t(2;2)(q35;p23)
PAX3-NCOA1
t(X;2)(q35;q13)
PAX3-AFX
+
Rhabdomyosarcoma
Embryonal
loss of heterozygosity 11p15,
trisomy 2, 8, 11, 12, 13 and 20
pathogenic variants
in RAS pathway genes
(NRAS, KRAS, HRAS, NF1, FGFR4)
+
Rhabdomyosarcoma
Sclerosing and spindle cell
VGLL2, TEAD1, SRF fusion
MYOD1 (p.L122R)
+
Synovial sarcomat(X;18)(p11,q11)
SS18-SSX1,
SS18-SSX2,
SS18-SSX4
+
Malignant peripheral nerve
sheath tumor
complex chromosomal aberrations,
pathogenic alterations
in SUZ12 and EED genes,
NF1 inactivation
+
Alveolar soft-part sarcomat(X;17)(p11;q25)
ASPL(ASPSCR1)-TFE3
+
Angiomatoid fibrous histiocytomat(12;16)(q13:p11)
FUS-ATF1
t(2;22)(q33;q12)
EWSR1-CREB1
t(12;22)(q13;q12)
EWSR1-ATF1
+
BCOR—rearranged sarcomainv(X)(p11.4p11.22)
BCOR-CCNB3
t(X;4)(p11;q31)
BCOR-MAML3
t(X;22)(p11;q13)
ZC3H7B-BCOR
+
CIC—rearranged sarcomat(4;19)(q35;q13)
t(10; 19)(q26;q13)
CIC-DUX4
t(X;19)(q13;q13.3)
CIC-FOXO4
+
Clear cell sarcomat(12;22)(q13;q12)
EWSR1-ATF1
t(2;22)(q33;q12)
EWSR1-CREB1
+
Dermatofibrosarcoma protuberanst(17;22)(q21;q13)
COL1A1-PDGFB,
ring chromosome
r(17;22)
+
Desmoid-type fibromatosis5q21 loss,
trisomy 8, 20,
pathogenic alterations in CTNNB1 gene
+
Desmoplastic small round
cell tumor
t(11;22)(p13;q12)
EWSR1-WT1
+
Dedifferentiated Liposarcomaring and marker chromosome,
12q13-15: MDM2, CDK4
region amplification
+
Epithelioid sarcomadeletion 22q
SMARCB1
t(8;22)(q22;q11)
t(10;22)
+
Epithelioid hemangioendotheliomat(1;3)(p36;q25),
WWTR1-CAMTA1,
t(X;11)(p11;q22)
YAP1-TFE3
+
Extraskeletal myxoid chondrosarcomat(9;22)(q22;q12)
EWSR1-NR4A3
t(9;17)(q22;q11)
TAF15 (TAF2N)-NR4A3
t(9;15)(q22;q21)
TCF12-NR4A3
t(3;9)(q11;q22)
TFG-NR4A3 fusion
t(9;17)(q22;q11)
RBP56-NR4A3 fusion
+
Giant cell fibroblastomat(17;22)(q22;q13)
COL1A1-PDGFB
+
Infantile fibrosarcomat(12;15)(p13;q25)
ETV6-NTRK3,
t(2;15)(p21;q25)
EML4-NTRK3,
LMNA-NTRK1,
1q deletion,
trisomy 8, 11, 17, 20
+
Inflammatory myofibroblastic tumorTranslocations involving
the 2p23 region;
fusions involving the ALK gene
(with multiple partner genes)
t(3;6)(q12;q22)
TFG-ROS1
+
LeiomyosarcomaComplex aberrations,
frequently with 1p deletion
+
Lipoblastomat(7;8) (q21q12)
COL1A2-PLAG1
del(8) (q12q24)
HAS2-PLAG1
t(8;14) (q12;q24)
PLAG1-RAD51L1
t(2;8) (q31;q12.1)
COL3A1-PLAG1
+
Low-grade fibromyxoid sarcomat(7;16)(q33;p11)
FUS-CREB3L2
t(11;16)(p11;p11)
FUS-CREB3L1
+
Mesenchymal chondrosarcomat(8;8)(q13;q21)
HEY1-NCOA2
+
Myoepitheliomat(6;22)(p21;q12)
EWSR1-POU5F1
t(1;22)(q23;q12)
EWSR1-PBX1
(19;22)(q13;q12)
EWSR1-ZNF444
+
Myxoid round cell liposarcomat(12;16)(q13;p11)
FUS-DDIT3
t(12;22)(q13;q12)
EWSR1-DDIT3 (CHOP)
+
Myxoinflammatory fibroblastic sarcomat(1;10)(p22;q24)
TGFBR3/MGEA5
+
Myxofibrosarcomaring chromosome+
Solitary fibrous tumorinv(12)(q13q13)
NAB2-STAT6
+
Undifferentiated embryonal sarcoma of the liver t(11;19)(q13,q13)
MALAT1-MHLB1
+
Table 6. Molecular markers of liver tumors.
Table 6. Molecular markers of liver tumors.
Tumor TypeGenes/Molecular Profiles Characteristically AlteredDiagnostic MarkerPrognostic, Predictive Markers
HepatoblastomaCTNNB1+-
APC+/--
NFE2L2-unfavorable
prognostic factor
Undifferentiated embryonal sarcoma
of the liver
(UES)
t(11;19)(q13;q13.4)+-
the C19MC
region
amplification
+-
Malignant rhabdoid tumor of the liverSMARCB1+-
Table 7. Molecular markers—retinoblastoma.
Table 7. Molecular markers—retinoblastoma.
Tumor TypeGenes/Molecular
Profiles Characteristically Altered
Diagnostic MarkerPrognostic, Predictive Markers
Retinoblastoma
RB1:
-
SNPs: exons 1–27 of the gene including assessment of splicing sites and mosaic type changes
-
CNV
-
methylation of the promoter region
+-
BCOR:
-
SNPs
-
fusions
-unfavorable
prognostic factor
MYCN (amplification)-unfavorable
prognostic factor
Table 8. Molecular markers—melanoma.
Table 8. Molecular markers—melanoma.
Tumor TypeGenes/Molecular
Profiles
Characteristically Altered
Diagnostic MarkerPrognostic, Predictive Markers
Spitzoid melanoma (SM)fusions involving ROS1, NTRK3, NTRK3, ALK, BRAF, MAPK, MET, RET genes+potential therapeutic targets
segmental rearrangements within chromosomes--
homozygous deletion of the 9p21 region+unfavorable prognostic factor
TERT (promoter changes-rare)-unfavorable prognostic factor
Melanoma arising
from a congenital melanocytic nevus (CMN)
NRAS (most commonly p.Q61K/R)+potential therapeutic targets
(depending on the alteration detected)
BRAF (most commonly p.V600E)-
TERT (promoter hypermethylation)+
segmental rearrangements within chromosomes-
Classic melanoma (“adult-type melanoma”)BRAF (most commonly p.V600E)+potential therapeutic targets
(depending on the alteration detected)
TERT (promoter changes)+
segmental rearrangements within chromosomes-
Table 9. Molecular markers—ovarian cancers.
Table 9. Molecular markers—ovarian cancers.
Tumor TypeGenes/Molecular Profiles Characteristically AlteredDiagnostic MarkerPrognostic, Predictive Markers
Sertoli–Leydig cell tumorsDICER1+-
Primary small cell carcinoma of the ovary, hypercalcemic type, SCCOHTSMARCA4+-
Table 10. Selected targeted treatments for pediatric solid tumours—authorized or under development.
Table 10. Selected targeted treatments for pediatric solid tumours—authorized or under development.
Specific Gene Mutation/AlterationTargeted TreatmentDevelopment Phase Clinical Trial IdentifierTarget Population
ALK alterations:
-
fusions
-
point mutations
AKL- inhibitors:Phase I/IINCT00939770Anaplastic lymphoma kinase (ALK) positive tumors,
relapsed or refractory solid tumors or anaplastic large cell lymphoma,
CrizotinibPhase II/IIINCT03874273inflamatory myofibroblastic tumor
Phase IIINCT03126916neuroblastoma
Phase IINCT02034981patients harboring an alteration on ALK, MET or ROS1
EnsartinibPhase IINCT03213652Relapsed or refractory advanced solid tumors, Non-Hodgkin lymphoma, or histiocytic disorders with ALK or ROS1 alterations
Anti-CD 20 antibodyRituximabAuthorized-Mature B cell Lymphoma
Anti-GD 2 antibodyDinutuximabAuthorized-Neuroblastoma
Anti-CD-30 antibodyBrentuximab
Vedotin
Phase IIINCT02166463
NCT01979536
Hodgkin Lymphoma
ALCL
BRAF alterations:
-
point mutations
(including p.V600E)
-
Fusions
(KIAA1549:BRAF)
DabrafenibPhase I/IINCT01677741Advanced BRAF V600 mutation-positive solid tumors
Dabrafenib + TrametinibPhase IINCT02684058BRAF V600 mutation positive low grade glioma or relapsed or refractory high grade glioma
VemurafenibPhase IINCT03220035Relapsed or refractory advanced solid tumors, Non-Hodgkin Lymphoma, or histiocytic disorders with BRAF V600 mutations,
Langerhans cell histiocytosis (LCH),
and other histiocytic disorders.
CobimetinibPhase IINCT04079179Refractory langerhans cell histiocytosis (LCH), and other histiocytic disorders.
TrametinibPhase IINCT03363217Pediatric neuro-oncology patients with refractory tumor and activation of the MAPK/ERK pathway
SelumetinibPhase IIINCT04576117Recurrent or progressive low-grade glioma
CDKN2A/B deletionPalbociclibPhase IINCT03526250Rb positive advanced solid tumors,
Non-Hodgkin Lymphoma, or histiocytic disorders with activating alterations in cell cycle genes
Ribociclib with EverolimusPhase INCT03387020Recurrent or refractory malignant brain tumors
EZH2 alterationsTazemetostatAuthorized-Epithelioid sarcoma ≥16 years
FGRF alterationsErdafitinibPhase IINCT03210714Patients with relapsed or refractory advanced solid tumors, Non-Hodgkin lymphoma, or histiocytic disorders with FGFR alterations
H3-3A, HIST1H3B,
HIST1H3BC
point mutation
PanobinostatPhase INCT02717455DIPG (H3K27M)
VorinostatPhase IINCT02035137Neuroblastoma
GD2 CART-cellPhase INCT03635632Relapsed or refractory neuroblastoma and other GD2 positive cancers
LSD1SeclidemstatPhase INCT03600649Ewing or Ewing-related sarcomas
MEK alterationsCobimetinibPhase I/IINCT02639546Gliomas, sarcomas, neuroblastoma, melanoma, MPNST, rhabdoid tumors, including atypical teratoid/rhabdoid tumor (AT/RT), NF1-associated tumors or RASopathy-associated tumors
SelumetinibPhase IIINCT04576117Recurrent or progressive low-grade glioma
Authorized-Plexiform neurofibroma
MET:
-
amplifications
-
fusions
VolitinibPhase INCT03598244Recurrent or refractory primary CNS tumors
mTOR pathway genes alterations, including TSC1, TSC2EverolimusAuthorized-Subependymal giant cell astrocytoma (SEGA)
TemsirolimusPhase IIINCT02567435Rhabdomyosarcoma
NTRK gene fusionsVitrakvi/Larotrectinib
Entrectinib
Authorized-Treatment of adult and paediatric patients with solid tumours that display a neurotrophic tyrosine receptor kinase
(NTRK) gene fusion
PARP alterationsOlaparibPhase INCT04236414Pediatric solid tumours
PD-1/PD-L1PembrolizumabAuthorized-R/R classic Hodgkin Lymphoma,
melanoma ≥12 years
IpilimumabAuthorized-
PembrolizumabPhase INCT02359565Recurrent, progressive, or refractory high-grade gliomas, diffuse intrinsic pontine gliomas, hypermutated brain tumors, ependymoma or medulloblastoma
NivolumabPhase IINCT03173950Medulloblastoma, ependymoma, choroid plexus tumors, atypical/malignant
meningioma
RET alterationsSelpercatinibAuthorized-Treatment of adults and adolescents 12 years and older with advanced RET-mutant medullary thyroid cancer (MTC)
ROS1 fusionsRepotrectinibPhase I/IINCT04094610Pediatric and young adult subjects harboring ALK, ROS1, or NTRK1/2/3
EntrectinibPhase I/IINCT02650401Locally advanced or metastatic solid or primary CNS tumors
EnsartinibPhase IINCT03213652Relapsed or refractory advanced solid tumors, non-hodgkin lymphoma, or histiocytic disorders with ALK or ROS1 alterations
SMARCB1
-
point mutations
-
CNVs
TazemetostatPhase INCT02601937Rhabdoid tumors, INI1-negative tumors
SMO alterationsVismodegibPhase IINCT01878617Medulloblastoma SHH subtype
NCT01601184
SonidegibPhase I/IINCT01125800Medulloblastoma, advanced pediatric solid potentially dependent on the Hedgehog-signaling pathway
VEGFR, PDGFR alterationsPazopanibPhase IINCT01956669Pediatric solid tumors
RegorafenibPhase IINCT02048371Selected sarcoma subtypes:
(Ewing sarcoma, rhabdomyosarcoma,
osteosarcoma)
VEGFR1, VEGFR3, FGFR3, FGFR4, PDGFRA alterationsLenvatinibPhase I/IINCT02432274Refractory or relapsed solid malignancies
Table 11. Selected genetic syndromes associated with childhood cancers.
Table 11. Selected genetic syndromes associated with childhood cancers.
SyndromeCancersGene/Chromosome Region
Li-Fraumenisarcomas, leukemias, brain cancers, hepatoblastomaTP53
Xeroderma pigmentosummelanomaXPA, XPC, DDB2, ERCC2
Neurofibromatosis type 1lymphomas, brain cancers, sarcomas, optic nerve gliomas, meningiomas, Wilms tumor, rhabdomyosarcomaNF1, SPRED1
Ataxia-telangiectasiaCNS, GI tumors, leukemiasATM
Bloom syndromeacute leukemia, GI cancersBLM
Fanconi anemiaacute leukemia, liver tumorsFANCA, FANCB, FANCC, PALB2
and others
Nijmegen syndrome leukemias, lymphomas, medulloblastoma, glioma, rhabdomyosarcomaNBN
Beckwith-Wiedemann syndromenephroblastoma, hepatoblastoma, rhabdomyosarcoma, gonadoblastomaCDKN1C/11p15
Chromosomal syndromes
(Down syndrome, Klinefelter syndrome)
leukemias, CNS tumorstrisomy 21, 47XXY
Familial retinoblastomaretinoblastomaRB1
Familial Wilms tumornephroblastomaWT1, WT2 and others
Familial polyposis colihepatoblastomaAPC, MUTYH
Cardiofaciocutaneous syndrome (CFC) acute lymphoblastic leukemia, rhabdomyosarcoma, hepatoblastoma, lymphomasBRAF, MAP2K1, MAP2K2, KRAS,
Noonan syndromeneuroblastoma, acute lymphatic leukemia, glioma, rhabdosarcomaPTPN11, RAF1, BRAF, SOS1, NRAS, CBL
Costello syndromerhabdosarcoma, neuroblastoma, fibrosarcoma HRAS
Sotos syndromeWilms tumor, neuroblastoma, hepatoblastomaNSD1
Von Hippel–Lindau syndromerenal tumors, CNS tumors-especially of the cerebellum, tumors of the adrenal glands, and tumors of the retina.VHL
Gorlin syndromemedulloblastomaPTCH1
Rubinstein–Taybi syndromemedulloblastoma, meningiomas, acute lymphatic leukemia, pheochromocytoma, rhabdomyosarcomaCREBBP
Turcot syndromemedulloblastoma, gliomasAPC
DICER syndromepleuropulmonary blastoma, nephroblastoma, renal and brain sarcomas, thyroid adenomas and carcinomas, gonadal tumorsDICER1
Multiple endocrine neoplasia type 1 and 2adenomas/carcinomas of the endocrine systemMEN1 and RET
Tuberous sclerosisbrain and kidney tumorsTSC1 i TSC2
Trisomy 18hepatoblastomatrisomy 18
Simpson–Golabi–Behmel syndrome type 1hepatoblastomaGPC3
Glycogen storage disorder type 1a, III, IV, VIhepatoblastomaG6PC, AGL, GBE1, PYGL
Tyrosinemia type 1hepatocellular carcinomaFAH
Table 12. Germline alterations in selected solid tumors in children.
Table 12. Germline alterations in selected solid tumors in children.
Tumor TypeGene (MIM Number)
AT/RTSMARCB1 (MIM 601607)
SMARCA4 (MIM 603254)
Choroid plexus carcinomaTP53 (MIM 191170)
Congenital melanocytic nevi MC1R (MIM 155555)
Familial melanomasCDKN2A (MIM 600160),
CDK4 (MIM 123829)
Glioma of the optic pathwayNF1 (MIM 613675)
HemangioblastomaVHL (MIM 608537)
Malignant nerve sheath tumorNF1 (MIM 613675),
TP53 (MIM 191170)
MedulloblastomaAPC (MIM 611731)
BRCA2 (MIM 600185)
MLH1 (MIM 120436)
MSH2 (MIM 609309)
MSH6 (MIM 600678)
PMS2 (MIM 600259)
PALB2 (MIM 610355)
PTCH1 (MIM 601309)
SUFU (MIM 607035)
SMOH (MIM 601500)
TP53 (MIM 191170)
CREBBP (MIM 600140)
GLI3 (MIM 175700)
MeningiomaNF2 (MIM 607379)
PTCH1 (MIM 601309)
PTEN (MIM 601728)
SMARCB1 (MIM 601607
SMARCE1 (MIM 603111)
SUFU (MIM 607035)
WRN (MIM 604611)
MEN1(MIM 613733)
PineoblastomaDICER1 (MIM 606241)
RB1 (MIM 614041)
SchwannomaNF2 (MIM 607379)
PRKAR1A (MIM 188830)
SchwannomatosisLZTR1 (MIM 600574)
SMARCB1 (MIM 601607)
Spinal cord ependymomaNF2 (MIM 607379)
Subependymal giant cell astrocytomaTSC1/TSC2 (MIM 605284/191092)
NeuroblastomaPHOX2B (MIM 603851)
ALK (MIM 105590)
HepatoblastomaAPC (MIM 611731),
uniparental disomy at 11p15.5
RetinoblastomaRB1 (MIM 614041)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Trubicka, J.; Grajkowska, W.; Dembowska-Bagińska, B. Molecular Markers of Pediatric Solid Tumors—Diagnosis, Optimizing Treatments, and Determining Susceptibility: Current State and Future Directions. Cells 2022, 11, 1238. https://doi.org/10.3390/cells11071238

AMA Style

Trubicka J, Grajkowska W, Dembowska-Bagińska B. Molecular Markers of Pediatric Solid Tumors—Diagnosis, Optimizing Treatments, and Determining Susceptibility: Current State and Future Directions. Cells. 2022; 11(7):1238. https://doi.org/10.3390/cells11071238

Chicago/Turabian Style

Trubicka, Joanna, Wiesława Grajkowska, and Bożenna Dembowska-Bagińska. 2022. "Molecular Markers of Pediatric Solid Tumors—Diagnosis, Optimizing Treatments, and Determining Susceptibility: Current State and Future Directions" Cells 11, no. 7: 1238. https://doi.org/10.3390/cells11071238

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