Next Article in Journal
Further Evidence That Defects in Main Thyroid Dysgenesis-Related Genes Are an Uncommon Etiology for Primary Congenital Hypothyroidism in Mexican Patients: Report of Rare Variants in FOXE1, NKX2-5 and TSHR
Previous Article in Journal
Anthropometric Development in Children: Possible Changes in Body Mass, Basal Metabolic Rate and Inflammatory Status
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Genetics in Neuroblastoma Prognosis

1
Unit of Medical Genetics, IRCCS Istituto Giannina Gaslini, 16147 Genova, Italy
2
IRCCS Istituto Giannina Gaslini, 16147 Genova, Italy
3
Department of Pediatric Surgery, IRCCS Istituto Giannina Gaslini, 16147 Genova, Italy
4
Dipartimento di Neuroscienze, Riabilitazione, Oftalmologia, Genetica e Scienze Materno-Infantili, University of Genova, 16132 Genova, Italy
5
Laboratory of Molecular Biology, IRCCS Istituto Giannina Gaslini, 16147 Genova, Italy
6
Department of Pathology, IRCCS Istituto Giannina Gaslini, 16147 Genova, Italy
*
Author to whom correspondence should be addressed.
Authors share senior authorship.
Children 2021, 8(6), 456; https://doi.org/10.3390/children8060456
Submission received: 20 April 2021 / Revised: 23 May 2021 / Accepted: 27 May 2021 / Published: 29 May 2021
(This article belongs to the Section Pediatric Surgery)

Abstract

:
In recent years, much research has been carried out to identify the biological and genetic characteristics of the neuroblastoma (NB) tumor in order to precisely define the prognostic subgroups for improving treatment stratification. This review will describe the major genetic features and the recent scientific advances, focusing on their impact on diagnosis, prognosis, and therapeutic solutions in NB clinical management.

1. Introduction

Neuroblastoma (NB) is a pediatric heterogeneous disease with a median age of 17 months at diagnosis, which can evolve with a benign course or fatal illness, with a natural history ranging from a benign course to a terminal illness [1,2,3].
NB derives from neural crest progenitor cells with an overall incidence of 1 case per 100,000 children [4]. NB can arise in adrenal glands or in sympathetic ganglia with metastatic sites in bone marrow, lymph nodes, bone, liver, and orbital sites [5,6]. A subset of NB can spontaneously regress without treatment, while NB with widespread metastasis denotes a refractory disease and grim outcome [5]. The prognosis of NB ranges from spontaneous regression to progression, metastasis, and death with 5-year overall survival of more than 90% in the low-risk group, and about 40–50% in high-risk (HR) group patients [7,8]. Although this disease represents 8% of all malignant childhood cancers, it is responsible for 15% of pediatric cancer-related deaths [9].
NB patients have been classified into four categories (very low-, low-, intermediate-, high-risk) based on the presence of seven biological and clinical factors as suggested by the International Risk Group (INRG) [10]. The risk of death for each NB patient is defined based on disease presentation, age at diagnosis, tumor histology, tumor ploidy, localized or metastatic disease, recurrent segmental chromosome copy number alterations, and amplification of the proto-oncogene MYCN [7,10]. The current multimodal treatment of HR-NB includes surgery, chemotherapy, radiation, myeloablative chemotherapy with stem cell rescue, biological targeting, and immunotherapy [11,12]. The overall goals are minimizing surgical complications and reducing chemotherapy toxicity.
Tumor genetic analysis is a key component for risk stratification and prognosis in NB [7]. The molecular classification of NB tumors is currently routinely performed because of the influence of genetic variations both on treatment and clinical outcome. Recently, Coronado et al. suggested that the hemizygous deletion of chromosome 11q represents a biomarker of response to therapy with the anti-GD2 antibody combined with immune checkpoint inhibitors in HR-NB [13].
Although most cases of NB are sporadic and patients do not have a genetically inherited predisposition, rare familial cases present a genetic etiology in an autosomal dominant manner. [1]
In addition to the abovementioned biological and clinical factors, many genetic features contribute to better define NB patient prognosis. Past scientific efforts and the latest advances, supported by the employment of new technologies, contributed to the definition of the current NB risk stratification upon which the therapeutic treatment applied depends. In the last few years, the association between specific genetic variants and the risk of NB occurrence was discovered [14]. Genomic association studies (GWAS) identified risk variants in the following genes: CASC15, BARD1, CHEK2, LMO1, LIN28B, AXIN2, BRCA1, TP53, SMARCA4, and CDK1NB [1,14,15,16,17,18,19,20].
The purpose of the present review is to report the genetic implications known to date in the pathogenesis of NB and their correlation with the prognosis of the disease.

2. First Evidences of Genetics in NB

The patient’s age, clinical, and tumor classifications are the main factors affecting NB prognosis. Besides these, biological molecular markers have been introduced to better define the risk groups and to predict prognosis and disease recurrence.

2.1. MYCN

Among the MYC family of cellular proto-oncogenes, which regulates the expression of specific genes during growth and differentiation, MYCN and c-MYC are deeply involved in NB, the former evidently associated with a poor outcome [5]. Immunohistochemical studies demonstrated that tumors with normal levels of MYCN showed high levels of c-MYC, suggesting a correlation with poor outcome as well [21]. MYCN amplification, characterizing 25–30% of primary NB tumors, is the strongest independent prognostic factor for NB, regardless age and clinical stage, and it was the first used in risk classification [10]. It is highly associated with advanced stages of disease, rapid progression, and a poor prognosis, mainly found in HR-NB patients, but even in infants and patients with a lower stage of the disease [22,23]. MYCN was identified in 1983, and its genomic locus is the distal short arm of chromosome 2 (2p24.3) [21]. It is amplified both in primary tumors and cell lines, and cytogenetically identified in extrachromosomal acentric double minutes chromosomes and in homogenously staining regions other than chromosome 2. Dual-color FISH probes containing MYCN and LAF (2q11) control probes were used to assess MYCN status (amplified, not amplified) as recommended by the International Neuroblastoma Risk Group Biology Committee [24]. When the gene is amplified, it displays even more than 100 copies per nucleus. The amplified genomic region containing MYCN is usually on the order of 500 to 1000 kb, and additional genes located on 2p24.3 are frequently coamplified, such as DDX1 in approximately 40–50% of NB and NAG [25,26,27] and ALK genes in about 10% of NBs. Moreover, ALK has also been validated as a direct target of MYCN-mediated transcriptional regulation [28].

2.2. Segmental Chromosome Alterations

Different conditions of ploidy have been described in NB, with whole-chromosome copy number variations usually associated with patients younger than 1 year of age and with an excellent outcome [29]. Finally, different segmental chromosomal alterations (SCA) with heavy prognostic impact have been identified and include loss of chromosomes 1p [30], 3p [31], 4p [30], 6q [32,33], and 11q [34], thought to express tumor suppressor genes, and gains of chromosomes 1q [35], 2p [36], and 17q [37], carrying supposed oncogenes. Gain of 17q has been reported in more than 50% of cases of neuroblastoma [37], and loss of 1p has been identified in one-third of cases [35]. Both gain of 17q and loss of 1p have been associated with MYCN amplification and a very poor prognostic outcome. On the contrary, loss of 11q, detectable in one-third of HR-NB cases, mainly in older patients, is inversely correlated with MYCN amplification but, nevertheless, is associated with a poor prognosis, too [35]. Indeed, NB tumors with a genomic profile characterized by any of these recurrent segmental alterations (SCA profile) are significantly associated with a high risk of relapse and a frequently poor outcome, especially when compared with NB tumors bearing only numerical alterations (NCA profile) [38]. NB patients bearing tumors with NCA have a favorable prognosis; nevertheless, a small percentage (10–15%) relapse locally or die. Recently, it has been reported that the loss of the whole chromosome X has a negative prognostic value for NB patients with an NCA genomic profile, and it was associated with a higher risk of relapse [39]. Patients affected by NB with a genomic profile displaying both SCA and NCA share the poor outcome of those with SCA only [8]. Other segmental aberrations (such as loss of 8p, 9p, 14q, 19p, 19q, and 22q, or gain of 5p, 12q, and others) are presently called atypical or not recurrent SCAs, and the risk of poor clinical outcome associated with these chromosomal alterations is not yet as well established as it is for the recurrent ones [40]. An SCA profile is associated with a higher risk of metastatic relapse among patients with localized disease, too, and among infants with localized unresectable or metastatic NB without MYCN amplification [41]. The accumulation of segmental chromosome alterations leads to NB progression, with particular regard to older children [42]. The high frequency of SCA-related to rare recurrent mutations in known protein-coding genes indicates that SCAs are driver events for this disease [40], and they have been incorporated into the patient risk stratification [38]. Since chromosomal breakpoints occur on several different loci, they probably reveal defects in DNA maintenance or repair mechanisms [42]; otherwise, an elevated amount of double-stranded DNA could break as a result of the specific cellular physiology of NB [43]. Recently, it has been shown that the distal 6q27 deletion characterizes a group of HR-NB patients with a particularly poor prognosis [32,33], suggesting that such an aberration can determine a negative clinical outcome due to the loss of function of the 14 genes mapping in such a region. The minimal 6q27 deletion contains three genes, SFT2D1, RPS6KA2, and FGFR1OP, that can promote an aggressive NB phenotype. SFT2D1 is involved in vesicles transport, RPS6KA2 has tumor suppressive functions in solid tumors, and the FGFR1OP gene encodes a centrosome protein for microtubules anchoring, and it is involved in both the proliferation and differentiation of erythroid lineage [44]. FGFR1OP haploinsufficiency can lead to abnormal erythropoiesis, and a reduced number of red blood cells has been observed in patients with NB [33]. Further studies are warranted to confirm the biological role of these genes in NB tumorigenesis.

2.3. PHOX2B and ALK

Familial NB is a quite rare event observed in about 1% of patients [45]. Linkage studies firstly identified missense or frameshift mutations in the paired-like homeobox 2B (PHOX2B) gene in a small group of familial NBs associated with Hirschsprung disease, congenital hypoventilation syndrome, and neurofibromatosis. PHOX2B is located at 4p13 and encodes a transcription factor essential in neurogenesis regulation. Germline mutations of PHOX2B occur in 6% of hereditary cases of NB [8,46,47], while in sporadic cases, they have been rarely observed [48].
Mutations in the anaplastic lymphoma kinase (ALK) gene, located in the 2p23 region, are the major factors leading to NB familial predisposition [49,50]. ALK is a receptor tyrosine kinase acting in neuronal differentiation [51], and it shows a high expression in the embryonic nervous system but will significantly decrease after birth. In NB, the somatically acquired genomic amplification and activating mutations of ALK occur in 2–3% and 8–10% of primary tumors, respectively [48,52,53], and play an important role in NB oncogenesis [54], evidently associated with a poor outcome [4,5]. Germline mutations in the tyrosine kinase domain of ALK are different from that of the somatic mutations in sporadic cases: of the three main hotspot mutations, only the R1275 mutation has been recurrently observed in familial cases [48,55]. With regard to the F1174 and F1245 mutations, they have not been reported in familial cases, suggesting a more aggressive behavior with respect to the R1275 one [56]. Mutations that cause a constitutive activation of ALK result in an oncogenic activity, affecting downstream signaling pathways, such as the RAS/MAPK and the RET ones, and inducing cell transformation [57].

2.4. TRK

Somatic chromosomal translocations involving the NTRK1, NTRK2, and NTRK3 genes occur in approximately 1% of tumors. The tropomyosin-related kinase (Trk) family includes three receptor tyrosine kinases (RTKs) named TrkA, TrkB, and TrkC that are regulated by neurotrophins, growth factors involved in neuronal development and function. Each receptor shows specific affinity for different ligands: TrkA binds NGF (nerve growth factor), TrkB recognizes BDNF (brain derived neurotrophic factor), and TrkC binds NT-3 (neurotrophin-3) [58,59].
The activation of the different receptors can lead to distinct cellular outcomes, including neuronal differentiation, cell survival, development, and synaptic plasticity [60,61,62]. The neurotrophins–Trks binding activates the downstream Ras/MAPK and PI3K/Akt-mTOR signaling pathways, central for cell proliferation and survival. Activation of TRK family receptors is implicated in both pediatric and adult tumors. Indeed, many childhood tumors are characterized by TRK fusions. The specific EVT6–NTRK3 fusion has been detected in more than 75% of cases of certain rare cancers such as cellular congenital mesoblastic nephroma, infantile fibrosarcoma, and secretory carcinoma of the breast [63,64,65]. In other pediatric cancers, such as in spitzoid melanomas and in high grade glioma, the TRK fusion incidence is lower (from 10% to 40%). NBs that spontaneously regress or show high differentiation levels are characterized by high expression of TrkA, suggesting its favorable prognostic impact, while in the majority of HR-NB NBs, the overexpression of TrkB is associated with an unfavorable outcome and aggressive tumor progression. BDNF/TrkB signaling promotes cell survival and angiogenesis. Indeed, inhibiting TrkB activity leads to cell apoptosis and increased chemosensitivity [66,67]. On the contrary, TrkC is overexpressed in favorable neuroblastomas. Given the significant oncogenic role of TRK fusion proteins, their inhibition could provide an effective therapeutic strategy for NB. Recently, clinical trials have reported the use of larotrectinib, a highly selective pan-TRK inhibitor, in children with a change in NTRK1, NTRK2, or NTRK3 genes, which demonstrated a good response.

2.5. Mutations in Genes Other Than MYCN with Prognostic Value in NB

Despite MYCN oncogene aberrations playing a major role in NB development, many studies highlighted the presence of genetic variations in other genes contributing to NB occurrence. LIN28B exerts a negative regulation of let-7 miRNAs, leading to increased MYCN protein levels, and its overexpression or amplification in HR-NB tumors has been associated with a poor prognostic outcome [68]. LIN28B can also promote RAS-related nuclear protein (RAN) expression through direct binding of RAN mRNA or by downregulating the Ran stabilizing protein RANBP2 through let-7 miRNA suppression [69]. Moreover, LIN28B contributes to metastases formation by increasing the invasive and migratory ability of tumor cells and, thus, confers the aggressive phenotype HR-NB tumors [70,71]. Consequently, targeting LIN28B provides a valuable therapeutic approach. Difluoromethylornithine (DFMO) inhibits ornithine decarboxylase (ODC), which indirectly affects the LIN28B-mediated biological axis. DFMO treatment induced LIN28B downregulation, inhibiting NB tumor progression by decreasing the glycolytic metabolic rate of NB tumor cells [72]. Moreover, BET bromodomain inhibitor JQ1, in combination with panobinostat, showed strong antitumor effects in NB: it blocked BRD3 and BRD4 activity, preventing LIN28B transcriptional activation [73].
A GWAS study identified specific variants of BARD1 and LMO1 associated with HR-NB disease [74]. Interestingly, while the full length BARD1 transcript encodes for a protein with tumor suppressive function, alternative splicing variants produce protein products with opposite oncogenic properties [75]. Metastatic NB cases show a higher expression of the BARD1β transcript, which enhances tumor growth and leads to the acquisition of the most aggressive traits of the disease [76].
LMO1 exerts an oncogenic role: genetic variants occurring in this gene were associated with higher susceptibility to developing aggressive NB tumors, and its overexpression, occurring through chromosomal duplication events, was associated with a poor prognostic outcome [77].
The majority of adult tumors are characterized by mutations in TP53 gene encoding for the p53 protein, which plays a pivotal role in cell cycle regulation. Thus, aberrations in p53 function promote tumor development [78]. On the contrary, pediatric tumors, including NB, do not display TP53 mutations, but p53 activity is altered by the aberrant expression of MDM2, which negatively regulates p53 [79]. It has been demonstrated that higher expression levels of MDM2 in MYCN nonamplified NB patients have a prognostically negative impact [80]. Recently, two genomic amplifications on chromosome 12 have been observed, involving the mapping sites of MDM2, CDK4, and FRS2 genes, and these aberrations were correlated with NB patients’ poor clinical outcome [81]. In vitro studies on MYCN-amplified NB cell lines showed that MDM2 is able to act independently from p53 and to interact with the MYCN transcript, increasing its stability and, thus, favoring its expression [82]. Therefore, the protumorigenic effects of MDM2 occur both in MYCN nonamplified tumors via the p53 pathway and in MYCN-amplified tumors in a p53-independent mechanism [83]. Considering its strong tumor-enhancing properties, MDM2 has become an attractive potential therapeutic target. Nutlin-3 was the first developed small molecule effective in antagonizing MDM2 activity and counteracting tumor growth [84]. Further studies led to the production of even more effective and specific second-generation MDM2 inhibitors. Among them, RG7388 was able to significantly reduce tumor burden in NB xenograft models [85] and, when combined with chemotherapeutic drugs commonly employed for NB treatment, it showed synergistic effects in inducing tumor cell apoptosis [86].

3. Brand New Emerging Genetic Implications in NB

Translational research in clinical oncology has been highly affected by the postgenome era, which has led to the application of new technologies including next generation sequencing, gene expression, and proteomics analysis. These methods allow us to obtain large data sets to better understand the molecular profile of tumors, with the possibility of characterizing tumor heterogeneity and defining targeted therapeutic approaches. Beside the development of novel technologies, translational research has seen an increasing interest in the study of liquid biopsies for monitoring the evolution of the disease and patient response to treatment. Both aspects characterized recent relevant studies on NB.

3.1. Exploring the Prognostic Value of Liquid Biopsy-Derived Markers in NB

Primary tumor tissue analysis is the golden standard for the molecular characterization of oncologic diseases. However, tissue sampling is highly limited by the invasiveness of the surgical procedure and by the quality of biological material that is not always adequate for biomolecular analyses. The need for novel analytical tools in clinical oncology has led to the investigation of liquid biopsies as a compelling, less invasive alternative for tumor analysis. In particular, HR-NB tumors often infiltrate adjacent structures [87] interfering with tissue sampling and, thus, making liquid biopsies an essential tool for studying tumor biology. Biological fluids are a valuable source of circulating messenger RNA (mRNA), tumor DNA (ctDNA), cell-free DNA (cfDNA), and circulating tumor cells (CTCs) that may provide biomarkers for tumor diagnosis and prognosis [88].

3.1.1. Liquid Biopsies to Unveil Minimal Residual Disease (MRD)

The concept of MRD refers to the persistence of cancerous cells after the conclusion of the chemotherapeutic treatment. These drug-resistant tumor cells are the main cause of relapse occurrence in NB patients, often resulting in fatal clinical outcome [5]. Therefore, the accurate detection of MRD is critical to improve NB patient prognosis. Residual NB tumor cells responsible for MRD can be often detected in peripheral blood (PB) and bone marrow (BM) [89], making liquid biopsies the optimal method for such analysis.
The first efforts in investigating blood samples of NB patients showed that high levels of circulating tyrosine hydroxylase (TH), PHOX2B, and doublecortin (DCX) mRNAs in peripheral blood (PB) and bone marrow (BM) samples at diagnosis are indicative of poor treatment response and worse clinical outcome [90]. These results were obtained through real-time quantitative PCR (RQ-PCR) analysis, which is the most sensitive method for MRD detection. Stutterheim et al. showed that the evaluation of two different panels of RQ-PCR markers can increase the sensitivity of MRD assessment. In particular, authors identified a panel for BM samples analysis (including PHOX2B, TH, DDC, CHRNA3, and GAP43) and a panel for PB samples (including PHOX2B, TH, DDC, DBH, and CHRNA3). The assessment of both RQ-PCR marker panels is crucial to obtain the highest sensitivity for a proper MRD detection [91].
A recent Japanese study demonstrated the high prognostic value of a 7 NB-mRNA signature (CRMP1, DBH, DDC, GAP43, ISL1, PHOX2B, and TH) evaluated with droplet digital PCR (ddPCR), a technique providing higher accuracy and reproducibility than RQ-PCR [92]. It was shown that ddPCR outperforms RQ-PCR in detecting the expression of the 7 NB-mRNA signature in BM and PB samples of HR-NB patients.
Besides mRNA signature analysis, the residual circulating tumor cells can be isolated from blood samples. It has been demonstrated that CTCs are present in BM samples of HR-NB patients both at diagnosis and at relapse. In particular, CTC genomic analysis provided comparable results to primary tumor tissue analysis, allowing for the proper detection of the aberrations occurring in genome coding regions [93].
These results demonstrate that liquid biopsies may be instrumental for NB diagnosis when primary tissue sampling is not feasible.

3.1.2. CfDNA and CtDNA Analyses for Understanding NB Clonal Evolution and Relapse Occurrence

The amount of cfDNA has been correlated with tumor burden in NB patients [94], and it provides important information for understanding tumor heterogeneity, a main hallmark of NB.
CtDNA analysis may serve as a surrogate for tissue biopsy to evaluate the genomic profile of NB tumors, and several studies have focused on the feasibility of detecting the genetic variants of the primary tumor in circulating nucleic acids. Successful results were obtained for the assessment of both MYCN [95] and ALK status by analyzing ctDNA. The sensitivity and specificity of the results were ensured by the application of droplet digital PCR technology (ddPCR), which confirmed the feasibility of determining the MYCN and ALK copy number profile from blood samples [96,97]. As the presence of the segmental chromosomal alterations strongly affects NB prognosis, the possibility to assess the NB genomic profile by analyzing cfDNA when tumor DNA is not available has been investigated. It has been shown that cfDNA in serum and plasma samples can be used to detect a 17q gain in NB patients [98]. Similarly, it has been demonstrated that shallow whole genome sequencing on cfDNA can identify the same chromosomal aberrations detected in primary tumor tissue and, thus, cfDNA can be considered as a reliable alternative source for the study of NB molecular features [99].
Recent studies showed that cfDNA analysis allows us to detect different cellular subclones at diagnosis, each carrying specific genomic aberrations, which can be responsible for treatment resistance or relapse occurrence [100]. Studying the spatial and temporal evolution of such clones through cfDNA analysis represents an important tool for disease monitoring and for developing targeted treatment strategies. Nevertheless, the low burden of recurrent mutations in NB [101] hinders the detection of NB specifically derived cfDNA and, thus, to overcome such a limitation, novel deep sequencing technologies are required. To this purpose, recent studies focused on the evaluation of the methylation status of cfDNA. Applebaum M. et al. demonstrated that the level of 5-hydroxymethylcytosine, a marker of active gene expression, is associated with disease burden and is indicative of treatment response and relapse occurrence [102]. Van Zogchel LMJ et al. [103] proposed the analysis of hypermethylation of RASSF1A (RASSF1Am) as a circulating biomarker for ctDNA detection, showing that RASSF1Am combined with the mRNA signature of BM analysis is able to better stratify patients with minimal residual disease [103].
The efforts of optimizing ctDNA analysis are leading to new insights in NB diagnosis and prognosis, but the stability of ctDNA and cfDNA is threatened by nucleases that can easily degrade circulating-free molecules: the short half-life of cfDNA can provide a “real-time” picture of disease status, but several different factors affecting the clearance rate should be considered [104].

3.1.3. Circulating Exosomes for NB Patient Response and Chemotherapy Resistance

Biological fluids also contain circulating exosomes, small vesicles involved in cellular communication, which provide a more stable intravesicular environment. The content of the exosomes probably depends on what stress the cell is experiencing, and it can change over time depending on the different stressors or stimuli to which the cell is subjected [105]. The discovery that the contents of exosomes can be transmitted from one cell to another by fusion supports the idea that exosomes are dynamic mediators of cellular communication [106]. All these studies have led to defining that exosomes are circulating small vesicles containing nucleic acids, miRNA, and proteins that are able to strongly affect cell behavior. The exosomes are released by most cell types, including tumor cells. Indeed, higher amounts of exosomes are secreted in pathological conditions. Exosomes cargo reflects the content of the tumor cells of origin, representing a valuable source of cancer biomarkers [107,108,109,110,111].
It has been demonstrated that exosomal-microRNA (exo-miR) derived from the plasma of HR-NB patients is predictive of treatment response [112]. In particular, exo-miR-29c, -342-3p, and let-7b downregulation correlated with a poor response to induction chemotherapy treatment. Moreover, the exo-miR expression profile is able to provide a chemoresistance index predicting sensitivity/resistance to specific chemotherapeutic drugs, allowing us to define targeted treatment approaches for maximal HR-NB patient response [112].
In vitro studies confirmed the role of exo-miR in chemoresistance, as reported by Challagundla et al. [113], who demonstrated the importance of the exosome mediated cross talk between NB cells and the tumor microenvironment (TME). In particular, it was shown that NB cells transfer miR-21 to monocytes via exosomes. MiR-21 leads to the upregulation of monocyte miR-155 expression in a Toll-like receptor-8-(TLR8)-dependent manner. MiR-155 is then uploaded in monocyte-derived exosomes and delivered back to NB cells, where it induces the downregulation of TERF1, a telomerase inhibitor, leading to increased telomerase activity and, thus, chemotherapy resistance.
Considering all this evidence highlighting the potential prognostic value of liquid biopsies in NB, the latest research studies [114] aim at optimizing the standard operating procedures (SOPs) for the study of circulating molecules and exosomes. The establishment of SOPs and the prospective validation of the results will be mandatory for the application of liquid biopsies assessment in a clinical setting for NB diagnosis, treatment, and monitoring.

3.1.4. Exosomal Double-Stranded DNA for the Analysis of Tumor Mutational Profile

Exosomes released by cancer cells contain about 20 times more exosomal-DNA (exo-DNA) than normal fibroblast-derived exosomes [107]. The exosomes isolated from tumor cells included exo-DNA that can be longer than 10 Kb [108]. It has recently been shown that the exo-DNA contained in the exosomes of NB patients can be useful for the analysis of the parental tumor mutational profile [115]. Degli Esposti et al. demonstrated, by whole exome sequencing, that NB-plasma-derived exosomes contain >10 Kb fragments of double-stranded tumor DNA. The exo-DNA showed genetic mutations in known NB oncogenes and tumor suppressor genes that were detected also in the tumor of origin. The most frequently analyzed somatic mutations in NB exo-DNA occurred in ALK, CHD5, SHANK2, PHOX2B, TERT, FGFR1, and BRAF genes. Furthermore, high genomic amplification of MYCN, TERT, and SHANK2 genes has been observed [115]. Exo-DNA of NB relapsed patients carried mutations in ALK, TP53, and RAS/MAP genes, suggesting that these somatic genetic variants may be responsible for acquired treatment resistance [115]. Notably, a considerably higher number of somatic mutations has been identified in exo-DNA at relapse than at diagnosis, supporting the theory of a clonal evolution. Certainly, circulating exosomes in the plasma of NB patients can serve as a source for the analysis of somatic mutations occurring in the primary tumor. Degli Esposti et al. [115] provided evidence for the analysis of a novel source of circulating DNA that represents an alternative to cfDNA, which can derive from cell lysis or apoptosis [96,99,101]. Therefore, the exo-DNA released by living cells better depicts the tumor dynamics and the most aggressive cell subpopulations. More specifically, the exo-DNA is stable and less fragmented than cfDNA because it is protected within an intravesicular environment and, thus, it represents a more reliable biological source to be analyzed in a clinical setting. Interestingly, exo-DNA also captures metastases features that cannot be detected by only analyzing needle biopsies of the primary tumor. In fact, somatic mutations that were not present in tumor DNA at the onset were instead identified in exo-DNA, in particular variants of ALK, ATRX, NF1, and TERT genes [115]. In this regard, exo-DNA can be instrumental for analyzing the great intratumoral spatial heterogeneity typical of NB [100], and for developing targeted therapeutic approaches. In the future, exo-DNA could provide a noninvasive method for diagnostic purposes, risk classification and patient stratification, and monitoring the response to therapy for NB patients, especially in those cases in which tumor tissue biopsy would not be feasible nor informative.

3.2. Telomerase Activity in Neuroblastoma

MYCN amplification is considered essential to evaluate the disease and to stratify treatment, but this is not sufficient to ensure an accurate prognostic risk grouping [116,117].
The identification of novel prognostic markers can improve not only the accuracy of risk assessment but also the definition of targets to develop new therapies. An emerging new marker is telomerase which, by its action on telomere maintenance, significantly contributes to cell immortalization, drug resistance, and tumorigenesis in almost 90% of human tumors, including peripheral neuroblastic tumors [118].
Telomeres are repetitive nucleotide sequences rich in guanine residues at the ends of chromosomes with the main function of maintaining genomic integrity. It is known that telomere dysfunction, mainly due to telomere capping disruption, promotes breakage-fusion-bridge (BFB) cycles, resulting in chromosome instability (CIN) [118]. Such events contribute to the structural and numerical aneuploidy observed in the majority of tumors, including NB [119]. It has been hypothesized that aneuploidy and CIN induce intratumor heterogeneity, a hallmark exploited by cancer cells to increase their adaptive potential and, thus, their survival [120,121]. It is also known that the absence of a replication system able to properly preserve telomere length in somatic cells leads to the progressive physiological shortening of telomeres, which is responsible for replicative senescence and apoptosis. To escape this critical condition, cancer cells promote the re-expression of telomerase reverse transcriptase (TERT), the catalytic subunit of telomerase [122,123]. Thus, telomere dysfunction is an impermanent process required for CIN-promoted tumor initiation, but telomerase reactivation is subsequently needed to ensure tumor progression [124].
To maintain telomeres length, tumor cells are also able to activate an alternative lengthening of telomeres (ALT) mechanism, based on homologous recombination. In particular, tumors developing from mesenchymal tissues are mainly characterized by ALT activity [125]. NB tumors are included in such a category, as they originate from the peripheral nervous system. It has been demonstrated that the coexistence of cancer cell subpopulations with different telomere length within NB is significantly associated with poor clinical outcome and disease progression in NB patients [126]. The study by Pezzolo et al. suggested that the coexpression of ALT mechanism and TERT, observed in 60% of the NB tumors analyzed, may play a major role in NB tumor progression. Importantly, telomere maintenance mechanisms are associated with poor clinical outcome regardless of clinical stage and their activation has been observed in HR-NB tumors but not in localized NB cases. It is known that TERT is a direct target of the MYCN oncogene and 40% of MYCN-amplified HR-NB cases are associated with MYCN-induced up-regulation of TERT [127]. Besides MYCN regulation, TERT overexpression can occur through other mechanisms, including gene amplification, genomic rearrangements, and mutations within the promoter region. The absence of TERT promoter mutations in NB tumor samples allows us to exclude this mechanism as the main one responsible for TERT overexpression in NB [128,129].
In the whole-genome sequencing analysis of neuroblastoma, recurrent genomic rearrangements of the catalytic subunit of TERT (5p15.33) and of the alpha thalassemia/mental retardation syndrome X-linked (ATRX) genes have been identified [127], in addition to MYCN amplifications, ALK, and PHOX2B. A subgroup of HR-NB tumors is characterized by the active ALT mechanism, whose function is negatively regulated by the ATRX gene [130]. Indeed, ATRX missense, nonsense, and frameshift mutations have been identified in NB tumors and associated with ALT activation and poor outcome in MYCN nonamplified NB tumors [131,132]
Moreover, ATRX deletions have been reported in 11% of HR NBs [133,134] and result in the loss of the nuclear ATRX protein and ALT mechanism activation. Recent evidence revealed that patients with TERT or ALT activation and harboring alterations in the RAS/p53 cellular pathway are identified as very HR cases, who are prone to relapse and associated with a very poor clinical outcome [135].
Telomerase maintenance is considered a powerful prognostic marker for NB patients and, thus, it provides a target for potential novel therapeutic treatments. A recent study tested the efficacy of the BET bromodomain inhibitor OTX015, targeting the BET bromodomain protein BRD4 that induces TERT expression, and carfilzomib, a proteasome inhibitor [136]. The results showed a strong synergistic effect of the two drugs in blocking TERT overexpression, inducing NB cell apoptosis in vitro, and drastically reducing tumor progression in murine models. These findings can encourage the development of the first clinical trial based on a combination of OTX015 and carfilzomib for patients harboring TERT-rearranged NB tumors.

3.3. Hypoxia

Hypoxia is a main hallmark of solid tumors, and it is represented by a decreased concentration of oxygen within tumor masses. This condition has been associated with aggressive cancer phenotypes characterized by higher metastatic potential and chemoresistance [137,138]. A pioneer study investigating the relationship between hypoxia and neuroblastoma has been published by Fardin P. et al., who first determined a hypoxia gene signature (NB-hypo) by analyzing the gene expression of different NB cell lines [139].
Gene expression profiling performed on primary NB tumors revealed that the identified NB-hypo could efficiently differentiate NB patients with good and poor clinical outcomes, showing for the first time that hypoxia is an independent prognostic factor for NB risk stratification [140]. Further studies refined the NB-hypo by reducing the hypoxia-associated gene expression signature to seven genes (NB-hop). The signature pointed out that the overexpression of the FAM162a, PDK1, PGK1, and MTFP1 genes and the downregulation of the ALDOC gene were associated with an unfavorable outcome for NB patients [141]. Importantly, the NB-hop was able to further stratify unfavorable subgroups of patients among low-risk cases without MYCN amplification and intermediate-risk patients with a stage 3, MYCN nonamplified disease who were older than 18 months [141]. These results were obtained in the most numerous cohort of NB patients analyzed, thanks to the application of a novel complex bioinformatics pipeline able to make the expression data of different datasets comparable [141]. Thus, hypoxia was confirmed as an independent prognostic factor predicting NB patient clinical outcome. Furthermore, the unfavorable NB-hop expression was strongly correlated with telomerase activation and immunosuppressive tumor microenvironment [141].
Considering all these findings, hypoxia represents a valuable therapeutic target in NB. Hypoxia-activated prodrugs (HAPs) are specifically designed to target the hypoxic areas of tumors, as these compounds are activated via oxidoreductase-mediated reduction, which is irreversible in the absence of oxygen [142]. Recent evidence showed that the administration of Evofosfamide, a drug belonging to the HAP class, produced cytotoxic effects both in vitro and in NB xenograft models, where a reduction in tumor growth was reported. The antitumor activity of Evofosfamide was amplified when used in combination with Topotecan [143].
The high specificity of HAPs in impairing hypoxic tumor cell viability can exert an important therapeutic activity in cancer patients who are properly stratified according to the hypoxic tumor status [142]. Thus, scientific efforts aiming at identifying hypoxic signatures able to properly stratify NB patients are mandatory for assessing the efficacy of HAPs.

4. Conclusions

In recent years, many advances have been made in NB genomics. Numerous studies have been published on recurrent somatic mutations and chromosomal abnormalities that contribute significantly to NB genesis, but their role in NB metastasis and response to therapy has not been fully understood yet. The broad range of phenotypical heterogeneity in NB, partly due to the diverse genetic features of this tumor, makes the identification of druggable genetic traits mandatory. Recurrent chromosomal rearrangements predict the prognosis of NB, but the genes mapping in such loci and involved in NB tumorigenesis have still to be completely defined. The present review summarized the main genetic features affecting NB prognosis studied in both primary tumor tissue and liquid biopsies (Table 1).
The advent of the genomic era has provided significant improvements in NB genetic characterization, and the possibility to analyze big datasets of patients has surely given essential results for better understanding the genetic features involved in NB tumorigenesis and progression. NB is a fatal pediatric tumor for 50% of HR-NB patients, whose cure largely depends on the possibility of studying the expression of biological markers of tumor progression and response to therapy. Thus, biological tumor profiling is mandatory and it is usually performed on primary tumor tissue. NB tumor samples have been essential to characterize many molecular features of NB tumors; however, a high percentage of NB patients, especially the high-risk class, are characterized either by tissue biopsies with insufficient content of neoplastic cells or by the absence of tumor tissue to be analyzed. Needle biopsy analysis is limited (i) by the difficulty of sampling an adequate amount of tissue, (ii) by the invasive nature of this method that hinders its repeatability, and (iii) by the impossibility to capture tumor heterogeneity. All these reasons justify the strong interest of the scientific community in alternative and more easily accessible biological material, such as the liquid biopsy. Recent scientific efforts in the study of liquid biopsies have revealed the importance of biological fluids for monitoring treatment response and for the identification of specific mutations in NB cellular subclones, undetectable in the primary tumor and responsible for relapse occurrence. Such mutations can serve as novel therapeutic targets for the development of targeted therapeutic treatment aiming at improving the survival of HR-NB patients.

Author Contributions

Conceptualization, K.M. and M.M.; writing—original draft preparation, M.L., M.O., A.P., M.M., and K.M.; writing—review and editing, M.L., M.O., A.P., G.M., F.Z., M.M., and K.M.; supervision, M.M. and K.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barr, E.K.; Applebaum, M.A. Genetic Predisposition to Neuroblastoma. Children 2018, 5, 119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Janoueix-Lerosey, I.; Lopez-Delisle, L.; Delattre, O.; Rohrer, H. The ALK receptor in sympathetic neuron development and neuroblastoma. Cell Tissue Res. 2018, 372, 325–337. [Google Scholar] [CrossRef]
  3. Brodeur, G.M. Spontaneous regression of neuroblastoma. Cell Tissue Res. 2018, 372, 277–286. [Google Scholar] [CrossRef] [PubMed]
  4. Swift, C.C.; Eklund, M.J.; Kraveka, J.M.; Alazraki, A.L. Updates in Diagnosis, Management, and Treatment of Neuroblastoma. Radiographics 2018, 38, 566–580. [Google Scholar] [CrossRef] [PubMed]
  5. Maris, J.M. Recent Advances in Neuroblastoma. N. Engl. J. Med. 2010, 362, 2202–2211. [Google Scholar] [CrossRef] [Green Version]
  6. Dubois, S.G.; Kalika, Y.; Lukens, J.N.; Brodeur, G.M.; Seeger, R.C.; Atkinson, J.B.; Haase, G.M.; Black, C.T.; Perez, C.; Shimada, H.; et al. Metastatic Sites in Stage IV and IVS Neuroblastoma Correlate With Age, Tumor Biology, and Survival. J. Pediatr. Hematol. 1999, 21, 181–189. [Google Scholar] [CrossRef]
  7. Bosse, K.R.; Maris, J.M. Advances in the translational genomics of neuroblastoma: From improving risk stratification and revealing novel biology to identifying actionable genomic alterations. Cancer 2016, 122, 20–33. [Google Scholar] [CrossRef] [PubMed]
  8. Matthay, K.K.; Maris, J.M.; Schleiermacher, G.; Nakagawara, A.; Mackall, C.L.; Diller, L.; Weiss, W. Neuroblastoma. Nat. Rev. Dis. Prim. 2016, 2, 16078. [Google Scholar] [CrossRef]
  9. Newman, E.A.; Nuchtern, J.G. Recent biologic and genetic advances in neuroblastoma: Implications for diagnostic, risk stratification, and treatment strategies. Semin. Pediatr. Surg. 2016, 25, 257–264. [Google Scholar] [CrossRef]
  10. Cohn, S.L.; Pearson, A.D.J.; London, W.B.; Monclair, T.; Ambros, P.F.; Brodeur, G.M.; Faldum, A.; Hero, B.; Iehara, T.; Machin, D.; et al. The international neuroblastoma risk group (INRG) classification system: An INRG task force report. J. Clin. Oncol. 2009, 27, 289–297. [Google Scholar] [CrossRef]
  11. Ladenstein, R.; Valteau-Couanet, D.; Brock, P.; Yaniv, I.; Castel, V.; Laureys, G.; Malis, J.; Papadakis, V.; Lacerda, A.; Ruud, E.; et al. Randomized Trial of Prophylactic Granulocyte Colony-Stimulating Factor During Rapid COJEC Induction in Pediatric Patients With High-Risk Neuroblastoma: The European HR-NBL1/SIOPEN Study. J. Clin. Oncol. 2010, 28, 3516–3524. [Google Scholar] [CrossRef]
  12. Wienke, J.; Dierselhuis, M.P.; Tytgat, G.A.; Künkele, A.; Nierkens, S.; Molenaar, J.J. The immune landscape of neuroblastoma: Challenges and opportunities for novel therapeutic strategies in pediatric oncology. Eur. J. Cancer 2021, 144, 123–150. [Google Scholar] [CrossRef]
  13. Coronado, E.; Yañez, Y.; Vidal, E.; Rubio, L.; Vera-Sempere, F.; Cañada-Martínez, A.J.; Panadero, J.; Cañete, A.; Ladenstein, R.; Castel, V.; et al. Intratumoral immunosuppression profiles in 11q-deleted neuroblastomas provide new potential therapeutic targets. Mol. Oncol. 2021, 15, 364–380. [Google Scholar] [CrossRef] [PubMed]
  14. Russell, M.R.; Penikis, A.; Oldridge, D.A.; Alvarez-Dominguez, J.R.; McDaniel, L.; Diamond, M.; Padovan, O.; Raman, P.; Li, Y.; Wei, J.S.; et al. CASC15-S Is a Tumor Suppressor lncRNA at the 6p22 Neuroblastoma Susceptibility Locus. Cancer Res. 2015, 75, 3155–3166. [Google Scholar] [CrossRef] [Green Version]
  15. Pandey, G.K.; Mitra, S.; Subhash, S.; Hertwig, F.; Kanduri, M.; Mishra, K.; Fransson, S.; Ganeshram, A.; Mondal, T.; Bandaru, S.; et al. The Risk-Associated Long Noncoding RNA NBAT-1 Controls Neuroblastoma Progression by Regulating Cell Proliferation and Neuronal Differentiation. Cancer Cell 2014, 26, 722–737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Bosse, K.R.; Diskin, S.J.; Cole, K.A.; Wood, A.C.; Schnepp, R.W.; Norris, G.; Nguyen, L.; Jagannathan, J.; Laquaglia, M.; Winter, C.; et al. Common Variation at BARD1 Results in the Expression of an Oncogenic Isoform That Influences Neuroblastoma Susceptibility and Oncogenicity. Cancer Res. 2012, 72, 2068–2078. [Google Scholar] [CrossRef] [Green Version]
  17. Oldridge, D.A.; Wood, A.C.; Weichert-Leahey, N.; Crimmins, I.; Sussman, R.; Winter, C.; McDaniel, L.D.; Diamond, M.; Hart, L.S.; Zhu, S.; et al. Genetic predisposition to neuroblastoma mediated by a LMO1 super-enhancer polymorphism. Nature 2015, 528, 418–421. [Google Scholar] [CrossRef] [Green Version]
  18. Diskin, S.J.; Capasso, M.; Schnepp, R.W.; Cole, K.A.; Attiyeh, E.F.; Hou, C.; Diamond, M.; Carpenter, E.L.; Winter, C.; Lee, H.; et al. Common variation at 6q16 within HACE1 and LIN28B influences susceptibility to neuroblastoma. Nat. Genet. 2012, 44, 1126–1130. [Google Scholar] [CrossRef]
  19. McDaniel, L.D.; Conkrite, K.L.; Chang, X.; Capasso, M.; Vaksman, Z.; Oldridge, D.A.; Zachariou, A.; Horn, M.; Diamond, M.; Hou, C.; et al. Common variants upstream of MLF1 at 3q25 and within CPZ at 4p16 associated with neuroblastoma. PLoS Genet. 2017, 13, e1006787. [Google Scholar] [CrossRef]
  20. Capasso, M.; McDaniel, L.D.; Cimmino, F.; Cirino, A.; Formicola, D.; Russell, M.R.; Raman, P.; Cole, K.A.; Diskin, S.J. The functional variant rs34330 of CDKN1B is associated with risk of neuroblastoma. J. Cell. Mol. Med. 2017, 21, 3224–3230. [Google Scholar] [CrossRef] [Green Version]
  21. Brodeur, G.M.; Seeger, R.C.; Schwab, M.; Varmus, H.E.; Bishop, J.M. Amplification of N-myc in untreated human neuroblastomas correlates with advanced disease stage. Science 1984, 224, 1121–1124. [Google Scholar] [CrossRef]
  22. Seeger, R.C.; Brodeur, G.M.; Sather, H.; Dalton, A.; Siegel, S.E.; Wong, K.Y.; Hammond, D. Association of Multiple Copies of the N-mycOncogene with Rapid Progression of Neuroblastomas. N. Engl. J. Med. 1985, 313, 1111–1116. [Google Scholar] [CrossRef] [PubMed]
  23. Mathew, P.; Valentine, M.B.; Bowman, L.C.; Rowe, S.T.; Nash, M.B.; Valentine, V.A.; Cohn, S.L.; Castleberry, R.P.; Brodeur, G.M.; Look, A.T. Detection of MYCN Gene Amplification in Neuroblastoma by Fluorescence In Situ Hybridization: A Pediatric Oncology Group Study. Neoplasia 2001, 3, 105–109. [Google Scholar] [CrossRef] [Green Version]
  24. Squire, J.A.; Thorner, P.S.; Weitzman, S.; Maggi, J.D.; Dirks, P.; Doyle, J.; Hale, M.; Godbout, R. Co-amplification of MYCN and a DEAD box gene (DDX1) in primary neuroblastoma. Oncogene 1995, 10, 1417–1422. [Google Scholar] [PubMed]
  25. Amler, L.C.; Schürmann, J.; Schwab, M. The DDX1 gene maps within 400 kbp 5′ to MYCN and is frequently coamplified in human neuroblastoma. Genes Chromosom. Cancer 1996, 15, 134–137. [Google Scholar] [CrossRef]
  26. Wimmer, K.; Zhu, C.; Lamb, B.J.; Kuick, R.; Ambros, P.F.; Kovar, H.; Thoraval, D.; Motyka, S.; Alberts, J.R.; Hanash, S.M. Co-amplification of a novel gene, NAG, with the N-myc gene in neuroblastoma. Oncogene 1999, 18, 233–238. [Google Scholar] [CrossRef] [Green Version]
  27. Javanmardi, N.; Fransson, S.; Djos, A.; Umapathy, G.; Östensson, M.; Milosevic, J.; Borenäs, M.; Hallberg, B.; Kogner, P.; Martinsson, T.; et al. Analysis of ALK, MYCN, and the ALK ligand ALKAL2 (FAM150B/AUGα) in neuroblastoma patient samples with chromosome arm 2p rearrangements. Genes Chromosom. Cancer 2020, 59, 50–57. [Google Scholar] [CrossRef]
  28. Hasan, K.; Nafady, A.; Takatori, A.; Kishida, S.; Ohira, M.; Suenaga, Y.; Hossain, S.; Akter, J.; Ogura, A.; Nakamura, Y.; et al. ALK is a MYCN target gene and regulates cell migration and invasion in neuroblastoma. Sci. Rep. 2013, 3, 3450. [Google Scholar] [CrossRef] [Green Version]
  29. Look, A.T.; Hayes, F.A.; Shuster, J.J.; Douglass, E.C.; Castleberry, R.P.; Bowman, L.C.; Smith, E.I.; Brodeur, G.M. Clinical relevance of tumor cell ploidy and N-myc gene amplification in childhood neuroblastoma: A Pediatric Oncology Group study. J. Clin. Oncol. 1991, 9, 581–591. [Google Scholar] [CrossRef]
  30. Caron, H.; Van Sluis, P.; De Kraker, J.; Bökkerink, J.; Egeler, R.; Laureys, G.; Slater, R.; Westerveld, A.; Voûte, M.; Versteeg, R. Allelic Loss of Chromosome 1p as a Predictor of Unfavorable Outcome in Patients with Neuroblastoma. N. Engl. J. Med. 1996, 334, 225–230. [Google Scholar] [CrossRef] [Green Version]
  31. Spitz, R.; Hero, B.; Ernestus, K.; Berthold, F. Deletions in chromosome arms 3p and 11q are new prognostic markers in localized and 4s neuroblastoma. Clin. Cancer Res. 2003, 9, 52–58. [Google Scholar]
  32. Depuydt, P.; Boeva, V.; Hocking, T.D.; Cannoodt, R.; Ambros, I.M.; Ambros, P.F.; Asgharzadeh, S.; Attiyeh, E.F.; Combaret, V.; Defferrari, R.; et al. Genomic Amplifications and Distal 6q Loss: Novel Markers for Poor Survival in High-risk Neuroblastoma Patients. J. Natl. Cancer Inst. 2018, 110, 1084–1093. [Google Scholar] [CrossRef] [PubMed]
  33. Ognibene, M.; Morini, M.; Garaventa, A.; Podestà, M.; Pezzolo, A. Identification of a minimal region of loss on chromosome 6q27 associated with poor survival of high-risk neuroblastoma patients. Cancer Biol. Ther. 2020, 21, 391–399. [Google Scholar] [CrossRef]
  34. Carén, H.; Kryh, H.; Nethander, M.; Sjöberg, R.-M.; Träger, C.; Nilsson, S.; Abrahamsson, J.; Kogner, P.; Martinsson, T. High-risk neuroblastoma tumors with 11q-deletion display a poor prognostic, chromosome instability phenotype with later onset. Proc. Natl. Acad. Sci. USA 2010, 107, 4323–4328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Gilbert, F.; Feder, M.; Balaban, G.; Brangman, D.; Lurie, D.K.; Podolsky, R.; Rinaldt, V.; Vinikoor, N.; Weisband, J. Hu-man neuroblastomas and abnormalities of chromosomes 1 and 17. Cancer Res. 1984, 44, 5444–5449. [Google Scholar] [PubMed]
  36. Szewczyk, K.; Wieczorek, A.; Młynarski, W.; Janczar, S.; Woszczyk, M.; Gamrot, Z.; Chaber, R.; Wysocki, M.; Pogorzała, M.; Bik-Multanowski, M.; et al. Unfavorable Outcome of Neuroblastoma in Patients With 2p Gain. Front. Oncol. 2019, 9, 1018. [Google Scholar] [CrossRef]
  37. Bown, N.; Cotterill, S.; Łastowska, M.; O’Neill, S.; Pearson, A.D.J.; Plantaz, D.; Meddeb, M.; Danglot, G.; Brinkschmidt, C.; Christiansen, H.; et al. Gain of Chromosome Arm 17q and Adverse Outcome in Patients with Neuroblastoma. N. Engl. J. Med. 1999, 340, 1954–1961. [Google Scholar] [CrossRef]
  38. Schleiermacher, G.; Michon, J.; Huon, I.; D’Enghien, C.D.; Klijanienko, J.; Brisse, H.; Ribeiro, A.P.B.; Mosseri, V.; Rubie, H.; Munzer, C.; et al. Chromosomal CGH identifies patients with a higher risk of relapse in neuroblastoma without MYCN amplification. Br. J. Cancer 2007, 97, 238–246. [Google Scholar] [CrossRef] [Green Version]
  39. Parodi, S.; Pistorio, A.; Erminio, G.; Ognibene, M.; Morini, M.; Garaventa, A.; Gigliotti, A.R.; Haupt, R.; Frassoni, F.; Pezzolo, A. Loss of whole chromosome X predicts prognosis of neuroblastoma patients with numerical genomic profile. Pediatr. Blood Cancer 2019, 66, e27635. [Google Scholar] [CrossRef] [PubMed]
  40. Janoueix-Lerosey, I.; Schleiermacher, G.; Michels, E.; Mosseri, V.; Ribeiro, A.; Lequin, D.; Vermeulen, J.; Couturier, J.; Peuchmaur, M.; Valent, A.; et al. Overall Genomic Pattern Is a Predictor of Outcome in Neuroblastoma. J. Clin. Oncol. 2009, 27, 1026–1033. [Google Scholar] [CrossRef] [PubMed]
  41. Schleiermacher, G.; Michon, J.; Ribeiro, A.; Pierron, G.; Mosseri, V.; Rubie, H.; Munzer, C.; Bénard, J.; Auger, N.; Combaret, V.; et al. Segmental chromosomal alterations lead to a higher risk of relapse in infants with MYCN-non-amplified localised unresectable/disseminated neuroblastoma (a SIOPEN collaborative study). Br. J. Cancer 2011, 105, 1940–1948. [Google Scholar] [CrossRef]
  42. Schleiermacher, G.; Janoueix-Lerosey, I.; Ribeiro, A.; Klijanienko, J.; Couturier, J.; Pierron, G.; Mosseri, V.; Valent, A.; Auger, N.; Plantaz, D.; et al. Accumulation of Segmental Alterations Determines Progression in Neuroblastoma. J. Clin. Oncol. 2010, 28, 3122–3130. [Google Scholar] [CrossRef]
  43. Cheung, N.-K.V.; Dyer, M.A. Neuroblastoma: Developmental biology, cancer genomics and immunotherapy. Nat. Rev. Cancer 2013, 13, 397–411. [Google Scholar] [CrossRef] [Green Version]
  44. Morandi, F.; Barco, S.; Stigliani, S.; Croce, M.; Persico, L.; Lagazio, C.; Scuderi, F.; Belli, M.L.; Montera, M.; Cangemi, G.; et al. Altered erythropoiesis and decreased number of erythrocytes in children with neuroblastoma. Oncotarget 2017, 8, 53194–53209. [Google Scholar] [CrossRef]
  45. Kushner, B.H.; Gilbert, F.; Helson, L. Familial neuroblastoma. Case reports, literature review, and etiologic considerations. Cancer 1986, 57, 1887–1893. [Google Scholar] [CrossRef]
  46. Trochet, D.; Bourdeaut, F.; Janoueix-Lerosey, I.; Deville, A.; de Pontual, L.; Schleiermacher, G.; Coze, C.; Philip, N.; Frébourg, T.; Munnich, A.; et al. Germline Mutations of the Paired–Like Homeobox 2B (PHOX2B) Gene in Neuroblastoma. Am. J. Hum. Genet. 2004, 74, 761–764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Mosse, Y.P.; Laudenslager, M.; Khazi, D.; Carlisle, A.J.; Winter, C.L.; Rappaport, E.; Maris, J.M. Germline PHOX2B Mutation in Hereditary Neuroblastoma. Am. J. Hum. Genet. 2004, 75, 727–730. [Google Scholar] [CrossRef] [Green Version]
  48. Mossé, Y.P.; Laudenslager, M.; Longo, L.; Cole, K.A.; Wood, A.; Attiyeh, E.F.; Laquaglia, M.J.; Sennett, R.; Lynch, J.E.; Perri, P.; et al. Identification of ALK as a major familial neuroblastoma predisposition gene. Nat. Cell Biol. 2008, 455, 930–935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Bourdeaut, F.; Ferrand, S.; Brugières, L.; Hilbert, M.; Ribeiro, A.; Lacroix, L.; Bénard, J.; Combaret, V.; Michon, J.; Valteau-Couanet, D.; et al. ALK germline mutations in patients with neuroblastoma: A rare and weakly penetrant syndrome. Eur. J. Hum. Genet. 2011, 20, 291–297. [Google Scholar] [CrossRef] [Green Version]
  50. Nakagawara, A.; Li, Y.; Izumi, H.; Muramori, K.; Inada, H.; Nishi, M. Neuroblastoma. Jpn. J. Clin. Oncol. 2018, 48, 214–241. [Google Scholar] [CrossRef] [Green Version]
  51. Motegi, A.; Fujimoto, J.; Kotani, M.; Sakuraba, H.; Yamamoto, T. ALK receptor tyrosine kinase promotes cell growth and neurite outgrowth. J. Cell Sci. 2004, 117, 3319–3329. [Google Scholar] [CrossRef] [Green Version]
  52. Chen, Y.; Takita, J.; Choi, Y.L.; Kato, M.; Ohira, M.; Sanada, M.; Wang, L.; Soda, M.; Kikuchi, A.; Igarashi, T.; et al. Onco-genic mutations of ALK kinase in neuroblastoma. Nature 2008, 455, 971–974. [Google Scholar] [CrossRef]
  53. George, R.E.; Sanda, T.; Hanna, M.; Fröhling, S.; Luther, W.; Zhang, J.; Ahn, Y.; Zhou, W.; London, W.B.; McGrady, P.; et al. Activating mutations in ALK provide a therapeutic target in neuroblastoma. Nat. Cell Biol. 2008, 455, 975–978. [Google Scholar] [CrossRef] [PubMed]
  54. Janoueix-Lerosey, I.; Lequin, D.; Brugières, L.; Ribeiro, A.; De Pontual, L.; Combaret, V.; Raynal, V.; Puisieux, A.; Schleiermacher, G.; Pierron, G.; et al. Somatic and germline activating mutations of the ALK kinase receptor in neuroblastoma. Nat. Cell Biol. 2008, 455, 967–970. [Google Scholar] [CrossRef]
  55. Carén, H.; Abel, F.; Kogner, P.; Martinsson, T. High incidence of DNA mutations and gene amplifications of the ALK gene in advanced sporadic neuroblastoma tumours. Biochem. J. 2008, 416, 153–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. de Pontual, L.; Kettaneh, D.; Gordon, C.T.; Oufadem, M.; Boddaert, N.; Lees, M.; Balu, L.; Lachassinne, E.; Petros, A.; Mollet, J.; et al. Germline gain-of-function mutations of ALK disrupt central nervous system development. Hum. Mutat. 2011, 32, 272–276. [Google Scholar] [CrossRef] [Green Version]
  57. Otto, T.; Horn, S.; Brockmann, M.; Eilers, U.; Schüttrumpf, L.; Popov, N.; Kenney, A.M.; Schulte, J.H.; Beijersbergen, R.; Christiansen, H.; et al. Stabilization of N-Myc Is a Critical Function of Aurora A in Human Neuroblastoma. Cancer Cell 2009, 15, 67–78. [Google Scholar] [CrossRef] [Green Version]
  58. Benedetti, M.; Levi, A.; Chao, M.V. Differential expression of nerve growth factor receptors leads to altered binding affinity and neurotrophin responsiveness. Proc. Natl. Acad. Sci. USA 1993, 90, 7859–7863. [Google Scholar] [CrossRef] [Green Version]
  59. Bibel, M.; Hoppe, E.; Barde, Y. Biochemical and functional interactions between the neurotrophin receptors trk and p75NTR. EMBO J. 1999, 18, 616–622. [Google Scholar] [CrossRef] [Green Version]
  60. Anderson, D.J. Cell fate determination in the peripheral nervous system: The sympathoadrenal progenitor. J. Neurobiol. 1993, 24, 185–198. [Google Scholar] [CrossRef]
  61. Patapoutian, A.; Reichardt, L.F. Trk receptors: Mediators of neurotrophin action. Curr. Opin. Neurobiol. 2001, 11, 272–280. [Google Scholar] [CrossRef]
  62. Lu, Y.; Christian, K.; Lu, B. BDNF: A key regulator for protein synthesis-dependent LTP and long-term memory? Neurobiol. Learn. Mem. 2008, 89, 312–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Sheng, W.-Q.; Hisaoka, M.; Okamoto, S.; Tanaka, A.; Meis-Kindblom, J.M.; Kindblom, L.-G.; Ishida, T.; Nojima, T.; Hashimoto, H. Congenital-Infantile Fibrosarcoma. Am. J. Clin. Pathol. 2001, 115, 348–355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Bourgeois, J.M.; Knezevich, S.R.; Mathers, J.A.; Sorensen, P.H.B. Molecular Detection of the ETV6-NTRK3 Gene Fusion Differentiates Congenital Fibrosarcoma From Other Childhood Spindle Cell Tumors. Am. J. Surg. Pathol. 2000, 24, 937–946. [Google Scholar] [CrossRef] [PubMed]
  65. Knezevich, S.R.; Garnett, M.J.; Pysher, T.J.; Beckwith, J.B.; Grundy, P.E.; Sorensen, P.H. ETV6-NTRK3 gene fusions and trisomy 11 establish a histogenetic link between mesoblastic nephroma and congenital fibrosarcoma. Cancer Res. 1998, 58, 5046–5048. [Google Scholar]
  66. Ho, R.; Eggert, A.; Hishiki, T.; Minturn, J.E.; Ikegaki, N.; Foster, P.; Camoratto, A.M.; Evans, A.E.; Brodeur, G.M. Re-sistance to chemotherapy mediated by TrkB in neuroblastomas. Cancer Res. 2002, 62, 6462–6466. [Google Scholar]
  67. Evans, A.E.; Kisselbach, K.D.; Liu, X.; Eggert, A.; Ikegaki, N.; Camoratto, A.M.; Dionne, C.; Brodeur, G. Effect of CEP-751 (KT-6587) on neuroblastoma xenografts expressing TrkB. Med. Pediatr. Oncol. 2001, 36, 181–184. [Google Scholar] [CrossRef]
  68. Molenaar, J.J.; Domingo-Fernández, R.; Ebus, M.E.; Lindner, S.; Koster, J.; Drabek, K.; Mestdagh, P.; Van Sluis, P.; Valentijn, L.J.; Van Nes, J.; et al. LIN28B induces neuroblastoma and enhances MYCN levels via let-7 suppression. Nat. Genet. 2012, 44, 1199–1206. [Google Scholar] [CrossRef]
  69. Schnepp, R.W.; Diskin, S.J. LIN28B: An orchestrator of oncogenic signaling in neuroblastoma. Cell Cycle 2016, 15, 772–774. [Google Scholar] [CrossRef] [Green Version]
  70. Chen, D.; Cox, J.; Annam, J.; Weingart, M.; Essien, G.; Rathi, K.S.; Rokita, J.L.; Khurana, P.; Cuya, S.M.; Bosse, K.R.; et al. LIN28B promotes neuroblastoma metastasis and regulates PDZ binding kinase. Neoplasia 2020, 22, 231–241. [Google Scholar] [CrossRef]
  71. Corallo, D.; Donadon, M.; Pantile, M.; Sidarovich, V.; Cocchi, S.; Ori, M.; De Sarlo, M.; Candiani, S.; Frasson, C.; Distel, M.; et al. LIN28B increases neural crest cell migration and leads to transformation of trunk sympathoadrenal precursors. Cell Death Differ. 2019, 27, 1225–1242. [Google Scholar] [CrossRef]
  72. Lozier, A.M.; Rich, M.E.; Grawe, A.P.; Peck, A.S.; Zhao, P.; Chang, A.T.-T.; Bond, J.P.; Sholler, G.S. Targeting ornithine decarboxylase reverses the LIN28/Let-7 axis and inhibits glycolytic metabolism in neuroblastoma. Oncotarget 2014, 6, 196–206. [Google Scholar] [CrossRef] [PubMed]
  73. Shahbazi, J.; Liu, P.Y.; Atmadibrata, B.; Bradner, J.E.; Marshall, G.M.; Lock, R.; Liu, T. The Bromodomain Inhibitor JQ1 and the Histone Deacetylase Inhibitor Panobinostat Synergistically Reduce N-Myc Expression and Induce Anticancer Effects. Clin. Cancer Res. 2016, 22, 2534–2544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Nguyễn, L.B.; Diskin, S.J.; Capasso, M.; Wang, K.; Diamond, M.A.; Glessner, J.; Kim, C.; Attiyeh, E.F.; Mosse, Y.P.; Cole, K.; et al. Phenotype Restricted Genome-Wide Association Study Using a Gene-Centric Approach Identifies Three Low-Risk Neuroblastoma Susceptibility Loci. PLoS Genet. 2011, 7, e1002026. [Google Scholar] [CrossRef] [PubMed]
  75. Cimmino, F.; Formicola, D.; Capasso, M. Dualistic Role of BARD1 in Cancer. Genes 2017, 8, 375. [Google Scholar] [CrossRef] [Green Version]
  76. Capasso, M.; Devoto, M.; Hou, C.; Asgharzadeh, S.; Glessner, J.T.; Attiyeh, E.F.; Mosse, Y.P.; Kim, C.; Diskin, S.J.; Cole, K.A.; et al. Common variations in BARD1 influence susceptibility to high-risk neuroblastoma. Nat. Genet. 2009, 41, 718–723. [Google Scholar] [CrossRef] [Green Version]
  77. Wang, K.; Diskin, S.J.; Zhang, H.; Attiyeh, E.F.; Winter, C.; Hou, C.; Schnepp, R.W.; Diamond, M.; Bosse, K.; Mayes, P.A.; et al. Integrative genomics identifies LMO1 as a neuroblastoma oncogene. Nat. Cell Biol. 2010, 469, 216–220. [Google Scholar] [CrossRef]
  78. Olivier, M.; Hollstein, M.; Hainaut, P. TP53 Mutations in Human Cancers: Origins, Consequences, and Clinical Use. Cold Spring Harb. Perspect. Biol. 2009, 2, a001008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Barone, G.; Tweddle, D.A.; Shohet, J.M.; Chesler, L.; Moreno, L.; Pearson, A.D.J.; Van Maerken, T. MDM2-p53 Interaction in Paediatric Solid Tumours: Preclinical Rationale, Biomarkers and Resistance. Curr. Drug Targets 2014, 15, 114–123. [Google Scholar] [CrossRef] [Green Version]
  80. Inomistova, M.V.; Svergun, N.M.; Khranovska, N.M.; Skachkova, O.V.; Gorbach, O.I.; Klymnyuk, G.I. Prognostic signif-icance of MDM2 gene expression in childhood neuroblastoma. Exp. Oncol. 2015, 37, 111–115. [Google Scholar] [CrossRef] [Green Version]
  81. Amoroso, L.; Ognibene, M.; Morini, M.; Conte, M.; Di Cataldo, A.; Tondo, A.; D’Angelo, P.; Castellano, A.; Garaventa, A.; Lasorsa, V.A.; et al. Genomic coamplification of CDK4/MDM2/FRS2 is associated with very poor prognosis and atypical clinical features in neuroblastoma patients. Genes Chromosom. Cancer 2019, 59, 277–285. [Google Scholar] [CrossRef]
  82. Gu, L.; Zhang, H.; He, J.; Li, J.; Huang, M.; Zhou, M. MDM2 regulates MYCN mRNA stabilization and translation in human neuroblastoma cells. Oncogene 2011, 31, 1342–1353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. He, J.; Gu, L.; Zhang, H.; Zhou, M. Crosstalk between MYCN and MDM2-p53 signal pathways regulates tumor cell growth and apoptosis in neuroblastoma. Cell Cycle 2011, 10, 2994–3002. [Google Scholar] [CrossRef] [Green Version]
  84. Van Maerken, T.; Ferdinande, L.; Taildeman, J.; Lambertz, I.; Yigit, N.; Vercruysse, L.; Rihani, A.; Michaelis, M.; Cinatl, J.; Cuvelier, C.A.; et al. Antitumor Activity of the Selective MDM2 Antagonist Nutlin-3 Against Chemoresistant Neuroblastoma With Wild-Type p53. J. Natl. Cancer Inst. 2009, 101, 1562–1574. [Google Scholar] [CrossRef]
  85. Lakoma, A.; Barbieri, E.; Agarwal, S.; Jackson, J.; Chen, Z.; Kim, Y.; McVay, M.; Shohet, J.M.; Kim, E.S. The MDM2 small-molecule inhibitor RG7388 leads to potent tumor inhibition in p53 wild-type neuroblastoma. Cell Death Discov. 2015, 1, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Chen, L.; Rousseau, R.F.; Middleton, S.A.; Nichols, G.L.; Newell, D.R.; Lunec, J.; Tweddle, D.A. Pre-clinical evaluation of the MDM2-p53 antagonist RG7388 alone and in combination with chemotherapy in neuroblastoma. Oncotarget 2015, 6, 10207–10221. [Google Scholar] [CrossRef] [Green Version]
  87. Tolbert, V.P.; Matthay, K.K. Neuroblastoma: Clinical and biological approach to risk stratification and treatment. Cell Tissue Res. 2018, 372, 195–209. [Google Scholar] [CrossRef] [PubMed]
  88. Pinzani, P.; D’Argenio, V.; Del Re, M.; Pellegrini, C.; Cucchiara, F.; Salvianti, F.; Galbiati, S. Updates on liquid biopsy: Current trends and future perspectives for clinical application in solid tumors. Clin. Chem. Lab. Med. 2021. [Google Scholar] [CrossRef] [PubMed]
  89. Uemura, S.; Ishida, T.; Thwin, K.K.M.; Yamamoto, N.; Tamura, A.; Kishimoto, K.; Hasegawa, D.; Kosaka, Y.; Nino, N.; Lin, K.S.; et al. Dynamics of Minimal Residual Disease in Neuroblastoma Patients. Front. Oncol. 2019, 9, 455. [Google Scholar] [CrossRef]
  90. Viprey, V.F.; Gregory, W.M.; Corrias, M.V.; Tchirkov, A.; Swerts, K.; Vicha, A.; Dallorso, S.; Brock, P.; Luksch, R.; Valteau-Couanet, D.; et al. Neuroblastoma mRNAs Predict Outcome in Children With Stage 4 Neuroblastoma: A European HR-NBL1/SIOPEN Study. J. Clin. Oncol. 2014, 32, 1074–1083. [Google Scholar] [CrossRef]
  91. Stutterheim, J.; Gerritsen, A.; Zappeij-Kannegieter, L.; Yalcin, B.; Dee, R.; Van Noesel, M.M.; Berthold, F.; Versteeg, R.; Caron, H.N.; Van Der Schoot, C.E.; et al. Detecting Minimal Residual Disease in Neuroblastoma: The Superiority of a Panel of Real-Time Quantitative PCR Markers. Clin. Chem. 2009, 55, 1316–1326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Thwin, K.K.; Ishida, T.; Uemura, S.; Yamamoto, N.; Lin, K.S.; Tamura, A.; Kozaki, A.; Saito, A.; Kishimoto, K.; Mori, T.; et al. Level of Seven Neuroblastoma-Associated mRNAs Detected by Droplet Digital PCR Is Associated with Tumor Relapse/Regrowth of High-Risk Neuroblastoma Patients. J. Mol. Diagn. 2020, 22, 236–246. [Google Scholar] [CrossRef] [Green Version]
  93. Abbasi, M.R.; Rifatbegovic, F.; Brunner, C.; Ladenstein, R.; Ambros, I.M.; Ambros, P.F. Bone marrows from neuroblastoma patients: An excellent source for tumor genome analyses. Mol. Oncol. 2014, 9, 545–554. [Google Scholar] [CrossRef]
  94. Wang, X.; Wang, L.; Su, Y.; Yue, Z.; Xing, T.; Zhao, W.; Zhao, Q.; Duan, C.; Huang, C.; Zhang, D.; et al. Plasma cell-free DNA quantification is highly correlated to tumor burden in children with neuroblastoma. Cancer Med. 2018, 7, 3022–3030. [Google Scholar] [CrossRef]
  95. Gotoh, T.; Hosoi, H.; Iehara, T.; Kuwahara, Y.; Osone, S.; Tsuchiya, K.; Ohira, M.; Nakagawara, A.; Kuroda, H.; Sugimoto, T. Prediction ofMYCNAmplification in Neuroblastoma Using Serum DNA and Real-Time Quantitative Polymerase Chain Reaction. J. Clin. Oncol. 2005, 23, 5205–5210. [Google Scholar] [CrossRef]
  96. Lodrini, M.; Sprüssel, A.; Astrahantseff, K.; Tiburtius, D.; Konschak, R.; Lode, H.; Fischer, M.; Keilholz, U.; Eggert, A.; Deubzer, H.E. Using droplet digital PCR to analyze MYCN and ALK copy number in plasma from patients with neuroblastoma. Oncotarget 2017, 8, 85234–85251. [Google Scholar] [CrossRef] [Green Version]
  97. Combaret, V.; Iacono, I.; Bellini, A.; Brejon, S.; Bernard, V.; Marabelle, A.; Coze, C.; Pierron, G.; Lapouble, E.; Schleiermacher, G.; et al. Detection of tumorALKstatus in neuroblastoma patients using peripheral blood. Cancer Med. 2015, 4, 540–550. [Google Scholar] [CrossRef]
  98. Combaret, V.; Bs, S.B.; Bs, I.I.; Schleiermacher, G.; Pierron, G.; Ribeiro, A.; Bergeron, C.; Marabelle, A.; Puisieux, A. Determination of 17q gain in patients with neuroblastoma by analysis of circulating DNA. Pediatr. Blood Cancer 2011, 56, 757–761. [Google Scholar] [CrossRef] [PubMed]
  99. Van Roy, N.; Van Der Linden, M.; Menten, B.; Dheedene, A.; Vandeputte, C.; Van Dorpe, J.; Laureys, G.; Renard, M.; Sante, T.; Lammens, T.; et al. Shallow Whole Genome Sequencing on Circulating Cell-Free DNA Allows Reliable Noninvasive Copy-Number Profiling in Neuroblastoma Patients. Clin. Cancer Res. 2017, 23, 6305–6314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Chicard, M.; Colmet-Daage, L.; Clement, N.; Danzon, A.; Bohec, M.; Bernard, V.; Baulande, S.; Bellini, A.; Eve, L.; Pierron, G.; et al. Whole-Exome Sequencing of Cell-Free DNA Reveals Temporo-spatial Heterogeneity and Identifies Treatment-Resistant Clones in Neuroblastoma. Clin. Cancer Res. 2018, 24, 939–949. [Google Scholar] [CrossRef] [Green Version]
  101. Pinto, N.R.; Applebaum, M.A.; Volchenboum, S.L.; Matthay, K.K.; London, W.B.; Ambros, P.F.; Nakagawara, A.; Berthold, F.; Schleiermacher, G.; Park, J.R.; et al. Advances in Risk Classification and Treatment Strategies for Neuroblastoma. J. Clin. Oncol. 2015, 33, 3008–3017. [Google Scholar] [CrossRef] [PubMed]
  102. Applebaum, M.A.; Barr, E.K.; Karpus, J.; West-Szymanski, D.; Oliva, M.; Sokol, E.A.; Zhang, S.; Zhang, Z.; Zhang, W.; Chlenksi, A.; et al. 5-hydroxymethylcytosine profiles in circulating cell-free DNA associate with disease burden in children with neuroblastoma. Clin. Cancer Res. 2019, 26, 1309–1317. [Google Scholar] [CrossRef] [PubMed]
  103. Van Zogchel, L.M.J.; Van Wezel, E.M.; Van Wijk, J.; Stutterheim, J.; Bruins, W.S.C.; Zappeij-Kannegieter, L.; Slager, T.J.E.; Schumacher-Kuckelkorn, R.; Verly, I.R.N.; Van Der Schoot, C.E.; et al. Hypermethylated RASSF1A as Circulating Tumor DNA Marker for Disease Monitoring in Neuroblastoma. JCO Precis. Oncol. 2020, 19, 291–306. [Google Scholar] [CrossRef] [PubMed]
  104. Kustanovich, A.; Schwartz, R.; Peretz, T.; Grinshpun, A. Life and death of circulating cell-free DNA. Cancer Biol. Ther. 2019, 20, 1057–1067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Gangoda, L.; Boukouris, S.; Liem, M.; Kalra, H.; Mathivanan, S. Extracellular vesicles including exosomes are mediators of signal transduction: Are they protective or pathogenic? Proteomics 2015, 15, 260–271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Théry, C.; Zitvogel, L.; Amigorena, S. Exosomes: Composition, biogenesis and function. Nat. Rev. Immunol. 2002, 2, 569–579. [Google Scholar] [CrossRef]
  107. Thakur, B.K.; Zhang, H.; Becker, A.; Matei, I.; Huang, Y.; Costa-Silva, B.; Zheng, Y.; Hoshino, A.; Brazier, H.; Xiang, J.; et al. Double-stranded DNA in exosomes: A novel biomarker in cancer detection. Cell Res. 2014, 24, 766–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Kahlert, C.; Melo, S.; Protopopov, A.; Tang, J.; Seth, S.; Koch, M.; Zhang, J.; Weitz, J.; Chin, L.; Futreal, A.; et al. Identification of Double-stranded Genomic DNA Spanning All Chromosomes with Mutated KRAS and p53 DNA in the Serum Exosomes of Patients with Pancreatic Cancer. J. Biol. Chem. 2014, 289, 3869–3875. [Google Scholar] [CrossRef] [Green Version]
  109. Takahashi, A.; Okada, R.; Nagao, K.; Kawamata, Y.; Hanyu, A.; Yoshimoto, S.; Takasugi, M.; Watanabe, S.; Kanemaki, M.T.; Obuse, C.; et al. Exosomes maintain cellular homeostasis by excreting harmful DNA from cells. Nat. Commun. 2017, 8, 1–16. [Google Scholar] [CrossRef] [Green Version]
  110. Colletti, M.; Petretto, A.; Galardi, A.; Di Paolo, V.; Tomao, L.; Lavarello, C.; Inglese, E.; Bruschi, M.; Lopez, A.A.; Pascucci, L.; et al. Proteomic Analysis of Neuroblastoma-Derived Exosomes: New Insights into a Metastatic Signature. Proteomics 2017, 17, 23–24. [Google Scholar] [CrossRef]
  111. Fonseka, P.; Liem, M.; Ozcitti, C.; Adda, C.G.; Ang, C.-S.; Mathivanan, S. Exosomes from N-Myc amplified neuroblastoma cells induce migration and confer chemoresistance to non-N-Myc amplified cells: Implications of intra-tumour heterogeneity. J. Extracell. Vesicles 2019, 8, 1597614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Morini, M.; Cangelosi, D.; Segalerba, D.; Marimpietri, D.; Raggi, F.; Castellano, A.; Fruci, D.; De Mora, J.F.; Cañete, A.; Yáñez, Y.; et al. Exosomal microRNAs from Longitudinal Liquid Biopsies for the Prediction of Response to Induction Chemotherapy in High-Risk Neuroblastoma Patients: A Proof of Concept SIOPEN Study. Cancers 2019, 11, 1476. [Google Scholar] [CrossRef] [Green Version]
  113. . Challagundla, K.B.; Wise, P.M.; Neviani, P.; Chava, H.; Murtadha, M.; Xu, T.; Kennedy, R.; Ivan, C.; Zhang, X.; Vannini, I.; et al. Exosome-Mediated Transfer of microRNAs Within the Tumor Microenvironment and Neuroblastoma Resistance to Chemotherapy. J. Natl. Cancer Inst. 2015, 107, 135. [Google Scholar] [CrossRef] [Green Version]
  114. Gerber, T.; Taschner-Mandl, S.; Saloberger-Sindhöringer, L.; Popitsch, N.; Heitzer, E.; Witt, V.; Geyeregger, R.; Hutter, C.; Schwentner, R.; Ambros, I.M.; et al. Assessment of Pre-Analytical Sample Handling Conditions for Comprehensive Liquid Biopsy Analysis. J. Mol. Diagn. 2020, 22, 1070–1086. [Google Scholar] [CrossRef]
  115. Degli Esposti, C.; Iadarola, B.; Maestri, S.; Beltrami, C.; Lavezzari, D.; Morini, M.; De Marco, P.; Erminio, G.; Garaventa, A.; Zara, F.; et al. Exosomes from Plasma of Neuroblastoma Patients Contain Double stranded DNA Reflecting the Mutational Status of Parental Tumor Cells. Int. J. Mol. Sci. 2021, 22, 3667. [Google Scholar] [CrossRef]
  116. Eggert, A.; Ikegaki, N.; Liu, X.-G.; Brodeur, G.M. Prognostic and Biological Role of Neurotrophin- Receptor TrkA and TrkB in Neuroblastoma1. Klin. Pädiatr. 2000, 212, 200–205. [Google Scholar] [CrossRef]
  117. Kawa, K.; Ohnuma, N.; Kaneko, M.; Yamamoto, K.; Etoh, T.; Mugishima, H.; Ohhira, M.; Yokoyama, J.; Bessho, F.; Honna, T.; et al. Long-Term Survivors of Advanced Neuroblastoma With MYCN Amplification: A Report of 19 Patients Surviving Disease-Free for More Than 66 Months. J. Clin. Oncol. 1999, 17, 3216–3220. [Google Scholar] [CrossRef]
  118. Bailey, S.M.; Murnane, J.P. Telomeres, chromosome instability and cancer. Nucleic Acids Res. 2006, 34, 2408–2417. [Google Scholar] [CrossRef] [Green Version]
  119. Tijhuis, A.E.; Johnson, S.C.; McClelland, S.E. The emerging links between chromosomal instability (CIN), metastasis, inflammation and tumour immunity. Mol. Cytogenet. 2019, 12, 1–21. [Google Scholar] [CrossRef] [Green Version]
  120. Nowell, P.C. The clonal evolution of tumor cell populations. Science 1976, 194, 23–28. [Google Scholar] [CrossRef]
  121. Targa, A.; Rancati, G. Cancer: A CINful evolution. Curr. Opin. Cell Biol. 2018, 52, 136–144. [Google Scholar] [CrossRef]
  122. Poremba, C.; Scheel, C.; Hero, B.; Christiansen, H.; Schaefer, K.-L.; Nakayama, J.-I.; Berthold, F.; Juergens, H.; Boecker, W.; Dockhorn-Dworniczak, B. Telomerase Activity and Telomerase Subunits Gene Expression Patterns in Neuroblastoma: A Molecular and Immunohistochemical Study Establishing Prognostic Tools for Fresh-Frozen and Paraffin-Embedded Tissues. J. Clin. Oncol. 2000, 18, 2582–2592. [Google Scholar] [CrossRef]
  123. Hiyama, E.; Hiyama, K.; Yokoyama, T.; Matsuura, Y.; Piatyszek, M.A.; Shay, J.W. Correlating telomerase activity levels with human neuroblastoma outcomes. Nat. Med. 1995, 1, 249–255. [Google Scholar] [CrossRef]
  124. Begus-Nahrmann, Y.; Hartmann, D.; Kraus, J.; Eshraghi, P.; Scheffold, A.; Grieb, M.; Rasche, V.; Schirmacher, P.; Lee, H.-W.; Kestler, H.A.; et al. Transient telomere dysfunction induces chromosomal instability and promotes carcinogenesis. J. Clin. Investig. 2012, 122, 2283–2288. [Google Scholar] [CrossRef]
  125. De Vitis, M.; Berardinelli, F.; Sgura, A. Telomere Length Maintenance in Cancer: At the Crossroad between Telomerase and Alternative Lengthening of Telomeres (ALT). Int. J. Mol. Sci. 2018, 19, 606. [Google Scholar] [CrossRef] [Green Version]
  126. Pezzolo, A.; Pistorio, A.; Gambini, C.; Haupt, R.; Ferraro, M.; Erminio, G.; De Bernardi, B.; Garaventa, A.; Pistoia, V. Intratumoral diversity of telomere length in individual neuroblastoma tumors. Oncotarget 2014, 6, 7493–7503. [Google Scholar] [CrossRef]
  127. Valentijn, L.J.; Koster, J.; Zwijnenburg, D.A.; Hasselt, N.E.; Van Sluis, P.; Volckmann, R.; Van Noesel, M.M.; George, R.E.; Tytgat, G.A.M.; Molenaar, J.J.; et al. TERT rearrangements are frequent in neuroblastoma and identify aggressive tumors. Nat. Genet. 2015, 47, 1411–1414. [Google Scholar] [CrossRef]
  128. Peifer, M.; Hertwig, F.; Roels, F.; Dreidax, D.; Gartlgruber, M.; Menon, R.; Krämer, A.; Roncaioli, J.L.; Sand, F.; Heuckmann, J.M.; et al. Telomerase activation by genomic rearrangements in high-risk neuroblastoma. Nat. Cell Biol. 2015, 526, 700–704. [Google Scholar] [CrossRef]
  129. Lindner, S.; Bachmann, H.S.; Odersky, A.; Schaefers, S.; Klein-Hitpass, L.; Hero, B.; Fischer, M.; Eggert, A.; Schramm, A.; Schulte, J.H. Absence of telomerase reverse transcriptase promoter mutations in neuroblastoma. Biomed. Rep. 2015, 3, 443–446. [Google Scholar] [CrossRef] [Green Version]
  130. Clynes, D.; Jelinska, C.; Xella, B.; Ayyub, H.; Scott, C.; Mitson, M.; Taylor, S.S.; Higgs, D.R.; Gibbons, R.J. Suppression of the alternative lengthening of telomere pathway by the chromatin remodelling factor ATRX. Nat. Commun. 2015, 6, 1–11. [Google Scholar] [CrossRef]
  131. Pugh, T.J.; Morozova, O.; Attiyeh, E.F.; Asgharzadeh, S.; Wei, J.S.; Auclair, D.; Carter, S.L.; Cibulskis, K.; Hanna, M.; Kiezun, A.; et al. The genetic landscape of high-risk neuroblastoma. Nat. Genet. 2013, 45, 279–284. [Google Scholar] [CrossRef] [Green Version]
  132. Lundberg, G.; Sehic, D.; Länsberg, J.-K.; Øra, I.; Frigyesi, A.; Castel, V.; Navarro, S.; Piqueras, M.; Martinsson, T.; Noguera, R.; et al. Alternative lengthening of telomeres-An enhanced chromosomal instability in aggressive non-MYCN amplified and telomere elongated neuroblastomas. Genes Chromosom. Cancer 2011, 50, 250–262. [Google Scholar] [CrossRef]
  133. Molenaar, J.J.; Koster, J.; Zwijnenburg, D.A.; Van Sluis, P.; Valentijn, L.J.; Van Der Ploeg, I.; Hamdi, M.; Van Nes, J.; Westerman, B.A.; Van Arkel, J.; et al. Sequencing of neuroblastoma identifies chromothripsis and defects in neuritogenesis genes. Nat. Cell Biol. 2012, 483, 589–593. [Google Scholar] [CrossRef]
  134. Heaphy, C.M.; De Wilde, R.F.; Jiao, Y.; Klein, A.P.; Edil, B.H.; Shi, C.; Bettegowda, C.; Rodriguez, F.J.; Eberhart, C.G.; Hebbar, S.; et al. Altered Telomeres in Tumors with ATRX and DAXX Mutations. Science 2011, 333, 425. [Google Scholar] [CrossRef] [Green Version]
  135. Ackermann, S.; Cartolano, M.; Hero, B.; Welte, A.; Kahlert, Y.; Roderwieser, A.; Bartenhagen, C.; Walter, E.; Gecht, J.; Kerschke, L.; et al. A mechanistic classification of clinical phenotypes in neuroblastoma. Science 2018, 362, 1165–1170. [Google Scholar] [CrossRef] [Green Version]
  136. Chen, J.; Nelson, C.; Wong, M.K.K.; Tee, A.E.; Liu, P.Y.; La, T.; Fletcher, J.I.; Kamili, A.; Mayoh, C.; Bartenhagen, C.; et al. Targeted Therapy of TERT-Rearranged Neuroblastoma with BET Bromodomain Inhibitor and Proteasome Inhibitor Combination Therapy. Clin. Cancer Res. 2021, 27, 1438–1451. [Google Scholar] [CrossRef]
  137. Brahimi-Horn, M.C.; Chiche, J.; Pouysségur, J. Hypoxia and cancer. J. Mol. Med. 2007, 85, 1301–1307. [Google Scholar] [CrossRef] [Green Version]
  138. Hussein, D.; Estlin, E.J.; Dive, C.; Makin, G.W. Chronic hypoxia promotes hypoxia-inducible factor-1α–dependent resistance to etoposide and vincristine in neuroblastoma cells. Mol. Cancer Ther. 2006, 5, 2241–2250. [Google Scholar] [CrossRef] [Green Version]
  139. Fardin, P.; Barla, A.; Mosci, S.; Rosasco, L.; Verri, A.; Varesio, L. The l1-l2 regularization framework unmasks the hypoxia signature hidden in the transcriptome of a set of heterogeneous neuroblastoma cell lines. BMC Genom. 2009, 10, 1–16. [Google Scholar] [CrossRef] [Green Version]
  140. Fardin, P.; Barla, A.; Mosci, S.; Rosasco, L.; Verri, A.; Versteeg, R.; Caron, H.N.; Molenaar, J.J.; Øra, I.; Eva, A.; et al. A biology-driven approach identifies the hypoxia gene signature as a predictor of the outcome of neuroblastoma patients. Mol. Cancer 2010, 9, 1–15. [Google Scholar] [CrossRef] [Green Version]
  141. Cangelosi, D.; Morini, M.; Zanardi, N.; Sementa, A.R.; Muselli, M.; Conte, M.; Garaventa, A.; Pfeffer, U.; Bosco, M.C.; Varesio, L.; et al. Hypoxia Predicts Poor Prognosis in Neuroblastoma Patients and Associates with Biological Mechanisms Involved in Telomerase Activation and Tumor Microenvironment Reprogramming. Cancers 2020, 12, 2343. [Google Scholar] [CrossRef]
  142. Spiegelberg, L.; Houben, R.; Niemans, R.; de Ruysscher, D.; Yaromina, A.; Theys, J.; Guise, C.P.; Smaill, J.B.; Patterson, A.V.; Lambin, P.; et al. Hypoxia-activated prodrugs and (lack of) clinical progress: The need for hypoxia-based biomarker patient selection in phase III clinical trials. Clin. Transl. Radiat. Oncol. 2019, 15, 62–69. [Google Scholar] [CrossRef] [Green Version]
  143. Zhang, L.; Marrano, P.; Wu, B.; Kumar, S.; Thorner, P.S.; Baruchel, S. Combined Antitumor Therapy with Metronomic Topotecan and Hypoxia-Activated Prodrug, Evofosfamide, in Neuroblastoma and Rhabdomyosarcoma Preclinical Models. Clin. Cancer Res. 2015, 22, 2697–2708. [Google Scholar] [CrossRef] [Green Version]
Table 1. Markers of unfavorable prognosis in NB.
Table 1. Markers of unfavorable prognosis in NB.
BiomarkerBiological SourceGenetic EventAlteration ResultReferences
MYCNPrimary tumor/cfDNA/exoDNAAmplificationOverexpression[5,21,22,23,95,96]
PHOX2BPrimary tumor/PB and BM/exoDNASomatic mutationsOverexpression[46,47,90,91,92]
ALKPrimary tumor/cfDNA/exoDNAAmplification and somatic mutationsOverexpression[48,49,50,52,53,54,55,56,57,97,115]
TRKPrimary tumorFusionOverexpression[66,67]
LIN28BPrimary tumorMYCN-mediated transcriptional activationOverexpression[68,69]
BARD1Primary tumorAlternative splicingAcquired oncogenic properties[74,75,76]
LMO1Primary tumorDuplicationOverexpression[74,77]
MDM2Primary tumorAmplificationOverexpression[79,80,81,82,83]
CDK4Primary tumorAmplificationOverexpression[81]
FRS2Primary tumorAmplificationOverexpression[81]
THPB and BM-Overexpression[90,91,92]
DCXPB and BM-Overexpression[90]
DDCPB and BM-Overexpression[91,92]
CHRNA3BM-Overexpression[91]
GAP43PB and BM-Overexpression[92]
DBHPB and BM-Overexpression[91,92]
CRMP1PB and BM-Overexpression[92]
ISL1PB and BM-Overexpression[92]
RASSF1APB and BMHypermethylation-[103]
CHD5exoDNASomatic mutations-[115]
SHANK2exoDNASomatic mutations-[115]
FGFR1exoDNASomatic mutations-[115]
BRAFexoDNASomatic mutations-[115]
TP53exoDNASomatic mutations-[115]
NF1exoDNASomatic mutations-[115]
TERTPrimary tumor/exoDNAAmplification and rearrangementsOverexpression[118,127]
ATRXPrimary tumor/exoDNARearrangements and somatic mutationsOverexpression[127,130,131,132]
FAM162aPrimary tumor-Overexpression[141]
PDK1Primary tumor-Overexpression[141]
PGK1Primary tumor-Overexpression[141]
MTFP1Primary tumor-Overexpression[141]
ALDOCPrimary tumor-Downregulation[141]
miR-29cExosomes-Downregulation[112]
Let-7bExosomes-Downregulation[112]
miR-342Exosomes-Downregulation[112]
The table shows the list of molecular markers that define NB prognosis. For each marker, the biological source, the type of genetic alteration, and the subsequent aberrant expression associated with unfavorable NB prognosis have been reported. Abbreviations: PB = peripheral blood; BM = bone marrow; exoDNA = exosomal DNA.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lerone, M.; Ognibene, M.; Pezzolo, A.; Martucciello, G.; Zara, F.; Morini, M.; Mazzocco, K. Molecular Genetics in Neuroblastoma Prognosis. Children 2021, 8, 456. https://doi.org/10.3390/children8060456

AMA Style

Lerone M, Ognibene M, Pezzolo A, Martucciello G, Zara F, Morini M, Mazzocco K. Molecular Genetics in Neuroblastoma Prognosis. Children. 2021; 8(6):456. https://doi.org/10.3390/children8060456

Chicago/Turabian Style

Lerone, Margherita, Marzia Ognibene, Annalisa Pezzolo, Giuseppe Martucciello, Federico Zara, Martina Morini, and Katia Mazzocco. 2021. "Molecular Genetics in Neuroblastoma Prognosis" Children 8, no. 6: 456. https://doi.org/10.3390/children8060456

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