Neuroblastoma Treatment (PDQ®): Treatment - Health Professional Information [NCI]

Skip to the navigation

This information is produced and provided by the National Cancer Institute (NCI). The information in this topic may have changed since it was written. For the most current information, contact the National Cancer Institute via the Internet web site at http://cancer.gov or call 1-800-4-CANCER.

General Information About Neuroblastoma

Dramatic improvements in survival have been achieved for children and adolescents with cancer.[1] Between 1975 and 2010, childhood cancer mortality decreased by more than 50%.[1,2,3] For neuroblastoma, the 5-year survival rate increased over the same time, from 86% to 95% for children younger than 1 year and from 34% to 68% for children aged 1 to 14 years.[2] Childhood and adolescent cancer survivors require close monitoring because cancer therapy side effects may persist or develop months or years after treatment. (Refer to the PDQ summary on Late Effects of Treatment for Childhood Cancer for specific information about the incidence, type, and monitoring of late effects in childhood and adolescent cancer survivors.)

Incidence and Epidemiology

Neuroblastoma is the most common extracranial solid tumor in childhood. More than 650 cases are diagnosed each year in North America.[4,5] The prevalence is about 1 case per 7,000 live births; the incidence is about 10.54 cases per 1 million per year in children younger than 15 years. About 37% are diagnosed as infants, and 90% are younger than 5 years at diagnosis, with a median age at diagnosis of 19 months.[6] The data on age at diagnosis show that this is a disease of infancy, with the highest rate of diagnosis in the first month of life.[4,5,6]

The incidence of neuroblastoma in black children is slightly lower than that in white children.[7] However, there are also racial differences in tumor biology, with African Americans more likely to have high-risk disease and fatal outcomes.[8,9]

Population-based studies of screening for infants with neuroblastoma have demonstrated that spontaneous regression of neuroblastoma without clinical detection in the first year of life is at least as prevalent as clinically detected neuroblastoma.[10,11,12]

Epidemiologic studies have shown that environmental or other exposures have not been unequivocally associated with increased or decreased incidence of neuroblastoma.[13]

Anatomy

Neuroblastoma originates in the adrenal medulla and paraspinal or periaortic regions where sympathetic nervous system tissue is present.



Drawing shows parts of the body where neuroblastoma may be found, including the paraspinal nerve tissue and the adrenal glands. Also shown are the spine and right and left kidney.

Figure 1. Neuroblastoma may be found in the adrenal glands and paraspinal nerve tissue from the neck to the pelvis.

Genetic Predisposition

Studies analyzing constitutional DNA in rare cohorts of familial neuroblastoma patients have provided insight into the complex genetic basis for tumor initiation. About 1% to 2% of patients with neuroblastoma have a family history of neuroblastoma. These children are, on average, younger (9 months at diagnosis) and have multifocal primary neuroblastoma (about 20%).

Several germline mutations have been associated with a genetic predisposition to neuroblastoma, including the following:

  • ALK gene mutation. The primary cause of familial neuroblastoma (about 75% of familial cases) is a germline mutation in the ALK (anaplastic lymphoma kinase) gene.[14] Somatic mutation in ALK is also seen in sporadic neuroblastoma. ALK is a tyrosine kinase receptor mutated in some lymphomas (refer to the Genomic and Biologic Features of Neuroblastoma section of this summary for more information).
  • PHOX2B gene mutation. Rarely, familial neuroblastoma may be associated with congenital central hypoventilation syndrome (Ondine curse), which is caused by a germline mutation of the PHOX2B gene.[15] Most PHOX2B mutations causing Ondine curse or Hirschsprung disease are polyalanine repeats and are not associated with familial neuroblastoma. However, germline loss-of-function PHOX2B mutations have been identified in rare patients with sporadic neuroblastoma and Ondine curse and/or Hirschsprung disease.[16] Aberration of PHOX2B has not been seen in patients with sporadic neuroblastoma without associated Ondine curse or Hirschsprung disease.
  • Germline deletion at the 1p36 or 11q14-23 locus. In case studies, germline deletion at the 1p36 or 11q14-23 locus has been associated with familial neuroblastoma, and the same deletions are found somatically in sporadic neuroblastoma.[17,18]

Sporadic neuroblastoma may also show a germline contribution, either with modest effect sizes for common polymorphic alleles or with greater effect sizes for rare pathogenic variants. As an example of the latter, rare germline variants of BARD1 have been identified in children with high-risk neuroblastoma.[19]

Genome-wide association studies have identified several common genomic variables (single nucleotide polymorphisms [SNPs]) with modest effect size that are associated with neuroblastoma. A subset of these SNPs is associated with susceptibility to high-risk neuroblastoma, including variants related to the following:

  • BARD1 (chromosome 2q35).[20]
  • LMO1 (chromosome 11p15).[21]
  • LIN28B (chromosome 6q16).[22]
  • HACE1 (chromosome 6q16).[22]
  • CASC15/NBAT-1 (chromosome 6p22).[23,24]

Other SNPs are associated with susceptibility to low-risk neuroblastoma.[25] One example that illustrates a mechanism by which SNPs may contribute to neuroblastoma risk is the polymorphism in the first intron of the oncogene LMO1 that forms a GATA transcription factor-binding site in an enhancer.[21,26] This risk allele is associated with high expression of LMO1 in aggressive neuroblastoma. LMO1 protein is necessary for growth of neuroblastoma in vitro and enhances growth of neuroblastoma cell lines with low LMO1 expression.

Genomic and Biologic Features of Neuroblastoma

Neuroblastoma can be subdivided into a biologically defined subset that has a very favorable prognosis (i.e., low-risk neuroblastoma) and another group that has a guarded prognosis (i.e., high-risk neuroblastoma). While neuroblastoma in infants with tumors that have favorable biology is highly curable, only 50% of children with high-risk neuroblastoma are alive at 5 years from diagnosis, at best.

Low-risk neuroblastoma is usually found in children younger than 18 months with limited extent of disease; the tumor has changes, usually increases, in the number of whole chromosomes in the neuroblastoma cell. Low-risk tumors are hyperdiploid when examined by flow cytometry.[27,28] In contrast, high-risk neuroblastoma generally occurs in children older than 18 months, is often metastatic to bone, and usually has segmental chromosome abnormalities. They are near diploid or near tetraploid by flow cytometric measurement.[27,28,29,30,31,32,33] High-risk tumors also show exonic mutations (refer to the Exonic mutations in neuroblastoma section of this summary for more information), but most high-risk tumors lack mutations in genes that are recurrently mutated. Compared with adult cancers, neuroblastomas show a low number of mutations per genome that affect protein sequence (10-20 per genome).[19]

Key genomic characteristics of high-risk neuroblastoma that are discussed below include the following:

  • Segmental chromosomal aberrations, including MYCN gene amplification.
  • Low rates of exonic mutations, with activating mutations in ALK being the most common recurring alteration.
  • Genomic alterations that promote telomere lengthening.

Segmental chromosomal aberrations (includingMYCNgene amplification)

Segmental chromosomal aberrations, found most frequently in 1p, 1q, 3p, 11q, 14q, and 17p (and MYCN amplification), are best detected by comparative genomic hybridization and are seen in almost all high-risk and/or stage 4 neuroblastomas.[29,30,31,32,33] Among all patients with neuroblastoma, a higher number of chromosome breakpoints correlated with the following, whether or not MYCN amplification was considered:[29,30,31,32,33][Level of evidence: 3iiD]

  • Advanced age at diagnosis.
  • Advanced stage of disease.
  • Higher risk of relapse.
  • Poorer outcome.

In a study of unresectable primary neuroblastomas without metastases in children older than 12 months, segmental chromosomal aberrations were found in most, and older children were more likely to have them and to have more of them per tumor cell. In children aged 12 to 18 months, the presence of segmental chromosomal aberrations had a significant effect on event-free survival (EFS) but not on overall survival (OS). However, in children older than 18 months, there was a significant difference in OS in children with segmental chromosomal aberrations versus children without segmental chromosomal aberrations (67% vs. 100%), regardless of the histologic prognosis.[33]

Segmental chromosomal aberrations are also predictive of recurrence in infants with localized unresectable or metastatic neuroblastoma without MYCN gene amplification.[27,28]

MYCN amplification (defined as more than 10 copies per diploid genome) is one of the most common segmental chromosomal aberrations, detected in 16% to 25% of tumors.[34] For high-risk neuroblastoma, 40% to 50% of cases show MYCN amplification.[35] In all stages of disease, amplification of the MYCN gene strongly predicts a poorer prognosis in both time to tumor progression and OS in almost all multivariate regression analyses of prognostic factors.[27,28] Within the localized MYCN-amplified cohort, ploidy status may further predict outcome.[36] However, patients with hyperdiploid tumors with any segmental chromosomal aberrations do relatively poorly.[29]

Most unfavorable clinical and pathobiological features are associated, to some degree, with MYCN amplification; in a multivariable logistic regression analysis of 7,102 International Neuroblastoma Risk Group patients, pooled segmental chromosomal aberrations and gain of 17q were the only poor prognostic features not associated with MYCN amplification. However, segmental chromosomal aberrations at 11q are almost mutually exclusive of MYCN amplification.

Exonic mutations in neuroblastoma

Multiple reports have documented that a minority of high-risk neuroblastomas have a small number of low-incidence, recurrently mutated genes. The most commonly mutated gene is ALK, which is mutated in approximately 10% of patients (see below). Other genes with even lower frequencies of mutation include ATRX, PTPN11, ARID1A, and ARID1B.[37,38,39,40,41,42,43] As shown in Figure 2, most neuroblastoma cases lack mutations in genes that are altered in a recurrent manner.



Chart showing the landscape of genetic variation in neuroblastoma.

Figure 2. Data tracks (rows) facilitate the comparison of clinical and genomic data across cases with neuroblastoma (columns). The data sources and sequencing technology used were whole-exome sequencing (WES) from whole-genome amplification (WGA) (light purple), WES from native DNA (dark purple), Illumina WGS (green), and Complete Genomics WGS (yellow). Striped blocks indicate cases analyzed using two approaches. The clinical variables included were gender (male, blue; female, pink) and age (brown spectrum). Copy number alterations indicates ploidy measured by flow cytometry (with hyperdiploid meaning DNA index >1) and clinically relevant copy number alterations derived from sequence data. Significantly mutated genes are those with statistically significant mutation counts given the background mutation rate, gene size, and expression in neuroblastoma. Germline indicates genes with significant numbers of germline ClinVar variants or loss-of-function cancer gene variants in our cohort. DNA repair indicates genes that may be associated with an increased mutation frequency in two apparently hypermutated tumors. Predicted effects of somatic mutations are color coded according to the legend. Reprinted by permission from Macmillan Publishers Ltd: Nature Genetics (Pugh TJ, Morozova O, Attiyeh EF, et al.: The genetic landscape of high-risk neuroblastoma. Nat Genet 45 (3): 279-84, 2013), copyright (2013).

ALK, the exonic mutation found most commonly in neuroblastoma, is a cell surface receptor tyrosine kinase, expressed at significant levels only in developing embryonic and neonatal brains. Germline mutations in ALK have been identified as the major cause of hereditary neuroblastoma. Somatically acquired ALK-activating mutations are also found as oncogenic drivers in neuroblastoma.[42]

The presence of an ALK mutation correlates with significantly poorer survival in high-risk and intermediate-risk neuroblastoma patients. ALK mutation was examined in 1,596 diagnostic neuroblastoma samples.[42]ALK tyrosine kinase domain mutations occurred in 8% of samples-at three hot spots and 13 minor sites-and correlated significantly with poorer survival in patients with high-risk and intermediate-risk neuroblastoma. ALK mutations were found in 10.9% of MYCN-amplified tumors versus 7.2% of those without MYCN amplification. ALK mutations occurred at the highest frequency (11%) in patients older than 10 years.[42] The frequency of ALK aberrations was 14% in the high-risk neuroblastoma group, 6% in the intermediate-risk neuroblastoma group, and 8% in the low-risk neuroblastoma group.

Small-molecule ALK kinase inhibitors such as crizotinib are being developed and tested in patients with recurrent and refractory neuroblastoma.[42] (Refer to the Treatment Options Under Clinical Evaluation for Recurrent or Refractory Neuroblastoma section in the PDQ summary on Neuroblastoma Treatment for more information about crizotinib clinical trials.)

Genomic evolution of exonic mutations

There are limited data regarding the genomic evolution of exonic mutations from diagnosis to relapse for neuroblastoma. Whole-genome sequencing was applied to 23 paired diagnostic and relapsed neuroblastomas to define somatic genetic alterations associated with relapse,[44] while a second study evaluated 16 paired diagnostic and relapsed specimens.[45] Both studies identified an increased number of mutations in the relapsed samples compared with the samples at diagnosis.

  • The first study found increased incidence of mutations in genes associated with RAS-MAPK signaling at relapse than at diagnosis, with 15 of 23 relapse samples containing somatic mutations in genes involved in this pathway and each mutation consistent with pathway activation.[44]

    In addition, three relapse samples showed structural alterations involving MAPK pathway genes consistent with pathway activation, so aberrations in this pathway were detected in 18 of 23 relapse samples (78%). Aberrations were found in ALK (n = 10), NF1 (n = 2), and one each in NRAS, KRAS, HRAS, BRAF, PTPN11, and FGFR1. Even with deep sequencing, 7 of the 18 alterations were not detectable in the primary tumor, highlighting the evolution of mutation presumably leading to relapse and the importance of genomic evaluations of tissues obtained at relapse.

  • In the second study, ALK mutations were not observed in either diagnostic or relapse specimens, but relapse-specific recurrent single-nucleotide variants were observed in 11 genes, including the putative CHD5 neuroblastoma tumor suppressor gene located at chromosome 1p36.[45]

In a study of 276 neuroblastoma samples of all stages and from patients of all ages, very deep (33,000X) sequencing of two amplified ALK mutational hot spots revealed 4.8% clonal mutations and an additional 5% subclonal mutations, suggesting that subclonal mutations are common.[46] Deep sequencing can reveal the presence of mutations in tiny subsets of tumor cells that may be able to survive during treatment and grow to constitute a relapse.

Genomic alterations promoting telomere lengthening

Lengthening of telomeres, the tips of chromosomes, promotes cell survival. Telomeres otherwise shorten with each cell replication, resulting eventually in the lack of a cell's ability to replicate. Low-risk neuroblastomas have little telomere lengthening activity. Aberrant genetic mechanisms for telomere lengthening have been identified for high-risk neuroblastoma.[37,38,47] Thus far, the following three mechanisms, which appear to be mutually exclusive, have been described:

  • Chromosomal rearrangements involving a chromosomal region at 5p15.33 proximal to the TERT gene, which encodes the catalytic unit of telomerase, occur in approximately 25% of high-risk neuroblastoma cases and are mutually exclusive with MYCN amplifications and ATRX mutations.[37,38] The rearrangements induce transcriptional upregulation of TERT by juxtaposing the TERT coding sequence with strong enhancer elements.
  • Another mechanism promoting TERT overexpression is MYCN amplification,[48] which is associated with approximately 40% to 50% of high-risk neuroblastomas.
  • The ATRX mutation or deletion is found in 10% to 20% of high-risk neuroblastomas, almost exclusively in older children,[39] and is associated with telomere lengthening by a different mechanism, termed alternative lengthening of telomeres.[39,47]

Additional biological factors associated with prognosis

MYC and MYCN expression

Immunostaining for MYC and MYCN proteins on 357 undifferentiated/poorly differentiated neuroblastomas has demonstrated that elevated MYC/MYCN protein expression is prognostically significant.[49] Sixty-eight tumors highly expressed MYCN protein, and 81 were MYCN amplified. Thirty-nine tumors expressed MYC highly and were mutually exclusive of high MYCN expression. Segmental chromosomal aberrations were not examined in this study, except for MYCN amplification.[49]

  • Patients with favorable-histology (FH) tumors without high MYC/MYCN expression had favorable survival (3-year EFS, 89.7% ± 5.5%; 3-year OS, 97% ± 3.2%).
  • Patients with undifferentiated or poorly differentiated histology tumors without MYC/MYCN expression had a 3-year EFS rate of 63.1% ± 13.6% and a 3-year OS rate of 83.5% ± 9.4%.
  • Three-year EFS rates in patients with MYCN amplification, high MYCN expression, and high MYC expression were 48.1% ± 11.5%, 46.2% ± 12%, and 43.4% ± 23.1%, respectively, and OS rates were 65.8% ± 11.1%, 63.2% ± 12.1%, and 63.5% ± 19.2%, respectively.
  • Further, when high expression of MYC and MYCN proteins were analyzed with other prognostic factors, including MYC/MYCN gene amplification, high MYC and MYCN protein expression was independent of other prognostic markers.

Most neuroblastomas with MYCN amplification in the International Neuroblastoma Pathology Classification system have unfavorable histology, but about 7% have FH. Of those with MYCN amplification and FH, most do not express MYCN, despite the gene being amplified, and have a more favorable prognosis than those that express MYCN.[50] Segmental chromosomal aberration at 11q is almost mutually exclusive of MYCN amplification.

Neurotrophin receptor kinases

Expression of neurotrophin receptor kinases and their ligands vary between high-risk and low-risk tumors. TrkA is found on low-risk tumors, and absence of its ligand NGF is postulated to lead to spontaneous tumor regression. In contrast, TrkB is found in high-risk tumors that also express its ligand, BDNF, which promotes neuroblastoma cell growth and survival.[51]

Immune system inhibition

Anti-GD2 antibodies, along with modulation of the immune system to enhance antineuroblastoma activity, are often used to help treat neuroblastoma. The anti-GD2 antibody (3F8), used for treating neuroblastoma exclusively at one institution, utilizes natural killer cells to kill the neuroblastoma cells. However, the natural killer cells can be inhibited by the interaction of HLA antigens and killer immunoglobulin receptor subtypes. Thus, the patient's immune system genes can help determine response to immunotherapy for neuroblastoma.[52,53] A report on the effects of immune system genes on response to dinutuximab, a commercially available anti-GD2 antibody, awaits publication.

Neuroblastoma Screening

Current data do not support neuroblastoma screening. Screening at the ages of 3 weeks, 6 months, or 1 year did not lead to reduction in the incidence of advanced-stage neuroblastoma with unfavorable biological characteristics in older children, nor did it reduce overall mortality from neuroblastoma.[11,12] No public health benefits have been shown from screening infants for neuroblastoma at these ages. (Refer to the PDQ summary on Neuroblastoma Screening for more information.)

Evidence (against neuroblastoma screening):

  1. A large population-based North American study, in which most infants in Quebec were screened at the ages of 3 weeks and 6 months, has shown that screening detects many neuroblastomas with favorable characteristics [10,11] that would never have been detected clinically, apparently because of spontaneous regression of the tumors.
  2. Another study of infants screened at the age of 1 year shows similar results.[12]

Clinical Presentation

The most common presentation of neuroblastoma is an abdominal mass. The most frequent signs and symptoms of neuroblastoma are caused by tumor mass and metastases. They include the following:

  • Proptosis and periorbital ecchymosis: Common in high-risk patients and arise from retrobulbar metastasis.
  • Abdominal distention: May occur with respiratory compromise in infants due to massive liver metastases.
  • Bone pain: Occurs in association with metastatic disease.
  • Pancytopenia: May result from extensive bone marrow metastasis.
  • Fever, hypertension, and anemia: Occasionally found in patients without metastasis.
  • Paralysis: Neuroblastoma originating in paraspinal ganglia may invade through neural foramina and compress the spinal cord extradurally. Immediate treatment is given for symptomatic spinal cord compression. (Refer to the Treatment of Spinal Cord Compression section of this summary for more information.)
  • Watery diarrhea: On rare occasions, children may have severe, watery diarrhea caused by the secretion of vasoactive intestinal peptide by the tumor, or they may have protein-losing enteropathy with intestinal lymphangiectasia.[54] Vasoactive intestinal peptide secretion may also occur upon chemotherapeutic treatment, and tumor resection reduces vasoactive intestinal peptide secretion.[55]
  • Presence of Horner syndrome: Horner syndrome is characterized by miosis, ptosis, and anhidrosis. It may be caused by neuroblastoma in the stellate ganglion, and children with Horner syndrome without other apparent cause are also examined for neuroblastoma and other tumors.[56]
  • Subcutaneous skin nodules: Neuroblastoma subcutaneous metastases often have bluish discoloration of the overlying skin and is usually seen only in infants.

The clinical characteristics of neuroblastoma in adolescents are similar to those observed in children. The only exception is that bone marrow involvement occurs less frequently in adolescents, and there is a greater frequency of metastases in unusual sites such as lung or brain.[57]

Opsoclonus/myoclonus syndrome

Paraneoplastic neurologic findings, including cerebellar ataxia or opsoclonus/myoclonus, occur rarely in children with neuroblastoma.[58] Opsoclonus/myoclonus syndrome can be associated with pervasive and permanent neurologic and cognitive deficits, including psychomotor retardation. Neurologic dysfunction is most often a presenting symptom but may arise long after removal of the tumor.[59,60,61]

Patients who present with opsoclonus/myoclonus syndrome often have neuroblastomas with favorable biological features and are likely to survive, though tumor-related deaths have been reported.[59]

The opsoclonus/myoclonus syndrome appears to be caused by an immunologic mechanism that is not yet fully defined.[59,62] The primary tumor is typically diffusely infiltrated with lymphocytes.[63]

Some patients may respond neurologically to removal of the neuroblastoma, but improvement may be slow and partial; symptomatic treatment is often necessary. Adrenocorticotropic hormone or corticosteroid treatment can be effective, but some patients do not respond to corticosteroids.[60,62] Other therapy with various drugs, plasmapheresis, intravenous gamma globulin, and rituximab have been reported to be effective in selected cases.[60,64,65,66] The long-term neurologic outcome may be superior in patients treated with chemotherapy, possibly because of its immunosuppressive effects.[58,64]

Diagnosis

Diagnostic evaluation of neuroblastoma includes the following:

  • Tumor imaging: Imaging of the primary tumor mass is generally accomplished by computed tomography or magnetic resonance imaging (MRI) with contrast. Paraspinal tumors that might threaten spinal cord compression are imaged using MRI. Metaiodobenzylguanidine (mIBG) scanning may also be used.[67,68]
  • Urine catecholamine metabolites: Urinary excretion of the catecholamine metabolites vanillylmandelic acid (VMA) and homovanillic acid (HVA) per milligram of excreted creatinine is measured before therapy. Collection of urine for 24 hours is not needed. If elevated, these markers can be used to determine the persistence of disease.

    Serum catecholamines are not routinely used in the diagnosis of neuroblastoma except in unusual circumstances.

  • Biopsy: Tumor tissue is often needed to obtain all the biological data required for risk-group assignment and subsequent treatment stratification in current Children's Oncology Group (COG) clinical trials. There is an absolute requirement for tissue biopsy to determine the International Neuroblastoma Pathology Classification (INPC). In the risk/treatment group assignment schema for COG studies, INPC has been used to determine treatment for patients with stage 3 disease, patients with stage 4S disease, and patients aged 18 months or younger with stage 4 disease. Additionally, a significant number of tumor cells are needed to determine MYCN copy number, DNA index, and the presence of segmental chromosomal aberrations.

    For patients older than 18 months with stage 4 disease, bone marrow with extensive tumor involvement combined with elevated catecholamine metabolites may be adequate for diagnosis and assigning risk/treatment group; however, INPC cannot be determined from tumor metastatic to bone marrow. Testing for MYCN amplification may be successfully performed on involved bone marrow if there is at least 30% tumor involvement.

    In rare cases, neuroblastoma may be discovered prenatally by fetal ultrasonography.[69] Management recommendations are evolving with regard to the need for immediate diagnostic biopsy in infants aged 6 months and younger with suspected neuroblastoma tumors that are likely to spontaneously regress. In a COG study of expectant observation of small adrenal masses in neonates, biopsy was not required for infants; 81% of patients avoided undergoing any surgery at all.[70] In a German clinical trial, 25 infants aged 3 months and younger with presumed localized neuroblastoma were observed without biopsy for periods of 1 to 18 months before biopsy or resection. There were no apparent ill effects from the delay.[71]

The diagnosis of neuroblastoma requires the involvement of pathologists who are familiar with childhood tumors. Some neuroblastomas cannot be differentiated morphologically, via conventional light microscopy with hematoxylin and eosin staining alone, from other small round blue cell tumors of childhood, such as lymphomas, primitive neuroectodermal tumors, and rhabdomyosarcomas. In such cases, immunohistochemical and cytogenetic analysis may be needed to diagnose a specific small round blue cell tumor.

The minimum criterion for a diagnosis of neuroblastoma, as established by international agreement, is that diagnosis must be based on one of the following:

  1. An unequivocal pathologic diagnosis made from tumor tissue by light microscopy (with or without immunohistology or electron microscopy).[72]
  2. The combination of bone marrow aspirate or trephine biopsy containing unequivocal tumor cells (e.g., syncytia or immunocytologically positive clumps of cells) and increased levels of urinary catecholamine metabolites.[72]

Prognostic Factors

Between 1975 and 2010, the 5-year survival rate for neuroblastoma in the United States increased from 86% to 95% for children younger than 1 year and increased from 34% to 68% for children aged 1 to 14 years.[2] The 5-year OS for all infants and children with neuroblastoma has increased from 46% when diagnosed between 1974 and 1989, to 71% when diagnosed between 1999 and 2005.[73] This single statistic can be misleading because of the extremely heterogeneous prognosis based on the neuroblastoma patient's age, stage, and biology. However, studies demonstrate a significant improvement in survival for high-risk patients diagnosed and treated between 2000 and 2010 compared with those diagnosed from 1990 to 1999.[74] (Refer to Table 1 for more information.)

The prognosis for patients with neuroblastoma is related to the following:[75,76,77,78]

  • Age at diagnosis.
  • Site of primary tumor.
  • Tumor histology.
  • Regional lymph node involvement (in children older than 1 year; however, this is controversial).
  • Response to treatment.
  • Biological features. (Refer to the Genomic and Biologic Features of Neuroblastoma section of this summary for more information.)

Some of these prognostic factors have been combined to create risk groups to help define treatment. (Refer to the International Neuroblastoma Risk Group Staging System section and the Children's Oncology Group Neuroblastoma Risk Grouping section of this summary for more information.)

Age at diagnosis

The effect of age at diagnosis on 5-year survival is profound. According to the 1975 to 2006 U.S. Surveillance, Epidemiology, and End Results (SEER) statistics, the 5-year survival stratified by age is as follows:[73]

  • Age younger than 1 year - 90%.
  • Age 1 to 4 years - 68%.
  • Age 5 to 9 years - 52%.
  • Age 10 to 14 years - 66%.

Children of any age with localized neuroblastoma and infants aged 18 months and younger with advanced disease and favorable disease characteristics have a high likelihood of long-term, disease-free survival (DFS).[79] The prognosis for fetal and neonatal neuroblastoma is similar to that for older infants with neuroblastoma and similar biological features.[80] Older children with advanced-stage disease, however, have a significantly decreased chance for cure, despite intensive therapy.

The effect of patient age on prognosis is strongly influenced by clinical and pathobiological factors, as evidenced by the following:

  • Since 2000, nonrandomized studies of low-risk and intermediate-risk patients have demonstrated that patient age has no effect on outcome of International Neuroblastoma Staging System (INSS) stage 1 or 2A disease. However, stage 2B patients younger than 18 months had a 5-year OS of 99% ± 1% versus 90% ± 4% for children aged 18 months and older.[81]
  • In the COG intermediate-risk study A3961 (NCT00003093) that included only MYCN non-amplified tumors, infants with INSS stage 3 tumors were compared with children with INSS stage 3 favorable-histology tumors. When INSS stage 3 infants with any histology were compared with stage 3 children with favorable histology, only EFS rates, not OS rates, were significantly different (3-year EFS, 95% ± 2 % vs. 87% ± 3 %; OS, 98% ± 1% vs. 99% ± 1%).[82]

In North American clinical trials reported in the 1990s, infants aged 1 year and younger had a cure rate higher than 80%, while older children had a cure rate of 50% to 70% with then-current, relatively intensive therapy.[83,84,85,86]

Survival of patients with INSS stage 4 disease is strongly dependent on age. Children younger than 18 months at diagnosis have a good chance of long-term survival (i.e., a 5-year DFS rate of 50%-80%),[87,88] with outcome particularly dependent on MYCN status, tumor cell ploidy, and the pattern of chromosomal aberrations (numerical chromosomal aberrations and segmental chromosomal aberrations). Hyperdiploidy and numerical chromosomal aberrations confer a favorable prognosis while diploidy and segmental chromosomal aberrations are associated with early treatment failure.[84,89] Infants aged 18 months and younger at diagnosis with INSS stage 4 neuroblastoma who do not have MYCN gene amplification are categorized as intermediate risk and have a 3-year EFS of 81% and OS of 93%.[6,82,90,91,92] Infants younger than 12 months with INSS stage 4 disease and MYCN amplification are categorized as high risk and have a 3-year EFS of 10%.[90]

Adolescents and young adults

Neuroblastoma has a worse long-term prognosis in adolescents older than 10 years or adults, regardless of stage or site. The disease is more indolent in older patients than in children.

Although adolescent and young adult patients have infrequent MYCN amplification (9% in patients aged 10-21 years), older children with advanced disease have a poor rate of survival. Tumors from the adolescent and young adult population commonly have segmental chromosomal aberrations, and ALK and ATRX mutations are much more frequent.[19,33,93]

The 5-year EFS rate is 32% for patients between the ages of 10 years and 21 years and the OS rate is 46%; for stage 4 disease, the 10-year EFS rate is 3% and the and OS rate is 5%.[94] Aggressive chemotherapy and surgery have been shown to achieve a minimal disease state in more than 50% of these patients.[57,95,96] Other modalities, such as local radiation therapy, autologous stem cell transplant, and the use of agents with confirmed activity, may improve the poor prognosis for adolescents and adults.[94,95,96]

Site of primary tumor

Clinical and biological features of neuroblastoma differ by primary tumor site. In a study of data on 8,389 patients entered in clinical trials and compiled by the International Risk Group Project, the following results were observed:[97]

  • Adrenal primary tumors were more likely than tumors originating in other sites to be associated with unfavorable prognostic features, including MYCN amplification, even after researchers controlled for age, stage, and histologic grade. Adrenal neuroblastomas were also associated with a higher incidence of stage 4 tumors, segmental chromosomal aberrations, diploidy, unfavorable INPC histology, age younger than 18 months, and elevated levels of lactate dehydrogenase (LDH) and ferritin. The relative risk of MYCN amplification compared with adrenal tumors was 0.7 in abdominal nonadrenal tumors and about 0.1 in nonabdominal paraspinal tumors.
  • Thoracic tumors were compared with nonthoracic tumors; after researchers controlled for age, stage, and histologic grade, results showed thoracic tumor patients had fewer deaths and recurrences (HR, 0.79; 95% confidence interval [CI], 0.67-0.92) and thoracic tumors had a lower incidence of MYCN amplification (adjusted OR, 0.20; 95% CI, 0.11-0.39).

Multifocal (multiple primaries) neuroblastoma occurs rarely, usually in infants, and generally has a good prognosis.[98] Familial neuroblastoma and germline ALK gene mutation should be considered in patients with multiple primary neuroblastomas.

Tumor histology

Neuroblastoma tumor histology has a significant impact on prognosis and risk group assignment (refer to the Cellular Classification of Neuroblastic Tumors section and Table 4 of this summary for more information).

Histologic characteristics considered prognostically favorable include the following:

  • Cellular differentiation/maturation. Higher degrees of neuroblastic maturation confer improved prognosis for stage 4 patients with segmental chromosome changes without MYCN amplification. Neuroblastoma tumors containing many differentiating cells, termed ganglioneuroblastoma, can have diffuse differentiation conferring a very favorable prognosis or can have nodules of undifferentiated cells whose histology, along with MYCN status, determine prognosis.[99,100]
  • Schwannian stroma.
  • Cystic neuroblastoma. About 25% of reported neuroblastomas diagnosed in the fetus and neonate are cystic; cystic neuroblastomas have lower stages and a higher incidence of favorable biology.[80]

High mitosis/karyorrhexis index is considered a prognostically unfavorable histologic characteristic, but its prognostic ability is age dependent.[101,102]

In a COG study (P9641 [NCT00003119]) investigating the effect of histology, among other factors, on outcome, 87% of 915 children with stage 1 and stage 2 neuroblastoma without MYCN amplification were treated with initial surgery and observation. Patients (13%) who had or were at risk of developing symptomatic disease, or who had less than 50% tumor resection at diagnosis, or who had unresectable progressive disease after surgery alone, were treated with chemotherapy and surgery. Those with favorable histologic features reported a 5-year EFS of 90% to 94% and OS of 99% to 100%, while those with unfavorable histology had an EFS of 80% to 86% and an OS of 89% to 93%.[81]

Regional lymph node involvement

According to the INSS, the presence of cancer in the regional lymph nodes on the same side of the body as the primary tumor has no effect on prognosis. However, when lymph nodes with metastatic neuroblastoma cross the midline and are on the opposite sides of the body from the primary tumor, the patient is upstaged (refer to the Stage Information for Neuroblastoma section of this summary for more information), and a poorer prognosis is conferred. In the COG P9641 (NCT00003119) low-risk study, stage 2b patients (those with tumor-containing lymph nodes on the same side of the body cavity as the tumor, but not on the opposite side of the cavity), but not stage 1 or 2a patients, had a poorer outcome with unfavorable histology (86% ± 5% vs. 99% ± 1%). The poorer outcome was predominantly in patients older than 18 months.[81]

Response to treatment

Response to treatment has been associated with outcome. In patients with high-risk disease, the persistence of neuroblastoma cells in bone marrow after induction chemotherapy, for example, is associated with a poor prognosis, which may be assessed by sensitive minimal residual disease techniques.[103,104,105] Similarly, the persistence of mIBG-avid tumor measured as Curie score (refer to the Curie score and SIOPEN score section of this summary for more information about Curie scoring) in two or more sites after completion of induction therapy predicts a poor prognosis.[106] A decrease in mitosis and an increase in histologic differentiation of the primary tumor are also prognostic.[107]

The accuracy of prognostication based on decrease in primary tumor size is less clear. In a study conducted by seven large international centers, 229 high-risk patients were treated in a variety of ways, including surgical removal of the primary tumor, radiation to the tumor bed, and, in most cases, antiGD2 antibody-enhanced immunotherapy. Primary tumor response was measured in three ways: as 30% or greater reduction in the longest dimension, 50% or greater reduction in tumor volume, or 65% or greater reduction in tumor volume (calculated from three tumor dimensions, a conventional radiologic technique). The measurements were performed at diagnosis and after induction chemotherapy before primary tumor resection. None of the methods of measuring primary tumor response at end of induction chemotherapy were predictive of survival.[108]

Spontaneous Regression of Neuroblastoma

The phenomenon of spontaneous regression has been well described in infants with neuroblastoma, especially in infants with the 4S pattern of metastatic spread.[109] (Refer to the Stage Information for Neuroblastoma section of this summary for more information.)

Spontaneous regression generally occurs only in tumors with the following features:[110]

  • Near triploid number of chromosomes.
  • No MYCN amplification.
  • No loss of chromosome 1p.

Additional features associated with spontaneous regression include the lack of telomerase expression,[111,112] the expression of Ha-ras,[113] and the expression of the neurotrophin receptor TrkA, a nerve growth factor receptor.[114]

Studies have suggested that selected infants who appear to have asymptomatic, small, low-stage adrenal neuroblastoma detected by screening or during prenatal or incidental ultrasound examination often have tumors that spontaneously regress and may be observed safely without surgical intervention or tissue diagnosis.[115,116,117]

Evidence (observation [spontaneous regression]):

  1. In a COG study, 83 highly selected infants younger than 6 months with stage 1 small adrenal masses as defined by imaging studies were observed without biopsy. Surgical intervention was reserved for those with growth or progression of the mass or increasing concentrations of urinary catecholamine metabolites.[70]
    • Eighty-one percent were spared surgery, and all were alive after 2 years of follow-up (refer to the Surgery subsection of this summary for more information).
  2. In a German clinical trial, spontaneous regression and/or lack of progression occurred in 44 of 93 asymptomatic infants aged 12 months or younger with stage 1, 2, or 3 tumors without MYCN amplification. All were observed after biopsy and partial or no resection.[71] In some cases, regression did not occur until more than 1 year after diagnosis.
  3. In neuroblastoma screening trials in Quebec and Germany, the incidence of neuroblastoma was twice that reported without screening, suggesting that many neuroblastomas are never noted and spontaneously regress.[10,11,12]

References:

  1. Childhood cancer by the ICCC. In: Howlader N, Noone AM, Krapcho M, et al., eds.: SEER Cancer Statistics Review, 1975-2010. Bethesda, Md: National Cancer Institute, 2013, Section 29. Also available online. Last accessed May 31, 2017.
  2. Smith MA, Altekruse SF, Adamson PC, et al.: Declining childhood and adolescent cancer mortality. Cancer 120 (16): 2497-506, 2014.
  3. Childhood cancer. In: Howlader N, Noone AM, Krapcho M, et al., eds.: SEER Cancer Statistics Review, 1975-2010. Bethesda, Md: National Cancer Institute, 2013, Section 28. Also available online. Last accessed January 27, 2017.
  4. Howlader N, Noone AM, Krapcho M, et al., eds.: SEER Cancer Statistics Review, 1975-2009 (Vintage 2009 Populations). Bethesda, Md: National Cancer Institute, 2012. Also available online. Last accessed February 7, 2017.
  5. Gurney JG, Ross JA, Wall DA, et al.: Infant cancer in the U.S.: histology-specific incidence and trends, 1973 to 1992. J Pediatr Hematol Oncol 19 (5): 428-32, 1997 Sep-Oct.
  6. London WB, Castleberry RP, Matthay KK, et al.: Evidence for an age cutoff greater than 365 days for neuroblastoma risk group stratification in the Children's Oncology Group. J Clin Oncol 23 (27): 6459-65, 2005.
  7. Ward E, DeSantis C, Robbins A, et al.: Childhood and adolescent cancer statistics, 2014. CA Cancer J Clin 64 (2): 83-103, 2014 Mar-Apr.
  8. Henderson TO, Bhatia S, Pinto N, et al.: Racial and ethnic disparities in risk and survival in children with neuroblastoma: a Children's Oncology Group study. J Clin Oncol 29 (1): 76-82, 2011.
  9. Latorre V, Diskin SJ, Diamond MA, et al.: Replication of neuroblastoma SNP association at the BARD1 locus in African-Americans. Cancer Epidemiol Biomarkers Prev 21 (4): 658-63, 2012.
  10. Takeuchi LA, Hachitanda Y, Woods WG, et al.: Screening for neuroblastoma in North America. Preliminary results of a pathology review from the Quebec Project. Cancer 76 (11): 2363-71, 1995.
  11. Woods WG, Gao RN, Shuster JJ, et al.: Screening of infants and mortality due to neuroblastoma. N Engl J Med 346 (14): 1041-6, 2002.
  12. Schilling FH, Spix C, Berthold F, et al.: Neuroblastoma screening at one year of age. N Engl J Med 346 (14): 1047-53, 2002.
  13. Heck JE, Ritz B, Hung RJ, et al.: The epidemiology of neuroblastoma: a review. Paediatr Perinat Epidemiol 23 (2): 125-43, 2009.
  14. Mossé YP, Laudenslager M, Longo L, et al.: Identification of ALK as a major familial neuroblastoma predisposition gene. Nature 455 (7215): 930-5, 2008.
  15. Mosse YP, Laudenslager M, Khazi D, et al.: Germline PHOX2B mutation in hereditary neuroblastoma. Am J Hum Genet 75 (4): 727-30, 2004.
  16. Raabe EH, Laudenslager M, Winter C, et al.: Prevalence and functional consequence of PHOX2B mutations in neuroblastoma. Oncogene 27 (4): 469-76, 2008.
  17. Satgé D, Moore SW, Stiller CA, et al.: Abnormal constitutional karyotypes in patients with neuroblastoma: a report of four new cases and review of 47 others in the literature. Cancer Genet Cytogenet 147 (2): 89-98, 2003.
  18. Mosse Y, Greshock J, King A, et al.: Identification and high-resolution mapping of a constitutional 11q deletion in an infant with multifocal neuroblastoma. Lancet Oncol 4 (12): 769-71, 2003.
  19. Pugh TJ, Morozova O, Attiyeh EF, et al.: The genetic landscape of high-risk neuroblastoma. Nat Genet 45 (3): 279-84, 2013.
  20. Bosse KR, Diskin SJ, Cole KA, et al.: Common variation at BARD1 results in the expression of an oncogenic isoform that influences neuroblastoma susceptibility and oncogenicity. Cancer Res 72 (8): 2068-78, 2012.
  21. Oldridge DA, Wood AC, Weichert-Leahey N, et al.: Genetic predisposition to neuroblastoma mediated by a LMO1 super-enhancer polymorphism. Nature 528 (7582): 418-21, 2015.
  22. Diskin SJ, Capasso M, Schnepp RW, et al.: Common variation at 6q16 within HACE1 and LIN28B influences susceptibility to neuroblastoma. Nat Genet 44 (10): 1126-30, 2012.
  23. Russell MR, Penikis A, Oldridge DA, et al.: CASC15-S Is a Tumor Suppressor lncRNA at the 6p22 Neuroblastoma Susceptibility Locus. Cancer Res 75 (15): 3155-66, 2015.
  24. Pandey GK, Mitra S, Subhash S, et al.: The risk-associated long noncoding RNA NBAT-1 controls neuroblastoma progression by regulating cell proliferation and neuronal differentiation. Cancer Cell 26 (5): 722-37, 2014.
  25. Nguyen le B, Diskin SJ, Capasso M, et al.: Phenotype restricted genome-wide association study using a gene-centric approach identifies three low-risk neuroblastoma susceptibility Loci. PLoS Genet 7 (3): e1002026, 2011.
  26. Wang K, Diskin SJ, Zhang H, et al.: Integrative genomics identifies LMO1 as a neuroblastoma oncogene. Nature 469 (7329): 216-20, 2011.
  27. Cohn SL, Pearson AD, London WB, et al.: The International Neuroblastoma Risk Group (INRG) classification system: an INRG Task Force report. J Clin Oncol 27 (2): 289-97, 2009.
  28. Schleiermacher G, Mosseri V, London WB, et al.: Segmental chromosomal alterations have prognostic impact in neuroblastoma: a report from the INRG project. Br J Cancer 107 (8): 1418-22, 2012.
  29. Janoueix-Lerosey I, Schleiermacher G, Michels E, et al.: Overall genomic pattern is a predictor of outcome in neuroblastoma. J Clin Oncol 27 (7): 1026-33, 2009.
  30. Schleiermacher G, Michon J, Ribeiro A, 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 105 (12): 1940-8, 2011.
  31. Carén H, Kryh H, Nethander M, et al.: High-risk neuroblastoma tumors with 11q-deletion display a poor prognostic, chromosome instability phenotype with later onset. Proc Natl Acad Sci U S A 107 (9): 4323-8, 2010.
  32. Schleiermacher G, Janoueix-Lerosey I, Ribeiro A, et al.: Accumulation of segmental alterations determines progression in neuroblastoma. J Clin Oncol 28 (19): 3122-30, 2010.
  33. Defferrari R, Mazzocco K, Ambros IM, et al.: Influence of segmental chromosome abnormalities on survival in children over the age of 12 months with unresectable localised peripheral neuroblastic tumours without MYCN amplification. Br J Cancer 112 (2): 290-5, 2015.
  34. Ambros PF, Ambros IM, Brodeur GM, et al.: International consensus for neuroblastoma molecular diagnostics: report from the International Neuroblastoma Risk Group (INRG) Biology Committee. Br J Cancer 100 (9): 1471-82, 2009.
  35. Kreissman SG, Seeger RC, Matthay KK, et al.: Purged versus non-purged peripheral blood stem-cell transplantation for high-risk neuroblastoma (COG A3973): a randomised phase 3 trial. Lancet Oncol 14 (10): 999-1008, 2013.
  36. Bagatell R, Beck-Popovic M, London WB, et al.: Significance of MYCN amplification in international neuroblastoma staging system stage 1 and 2 neuroblastoma: a report from the International Neuroblastoma Risk Group database. J Clin Oncol 27 (3): 365-70, 2009.
  37. Peifer M, Hertwig F, Roels F, et al.: Telomerase activation by genomic rearrangements in high-risk neuroblastoma. Nature 526 (7575): 700-4, 2015.
  38. Valentijn LJ, Koster J, Zwijnenburg DA, et al.: TERT rearrangements are frequent in neuroblastoma and identify aggressive tumors. Nat Genet 47 (12): 1411-4, 2015.
  39. Cheung NK, Zhang J, Lu C, et al.: Association of age at diagnosis and genetic mutations in patients with neuroblastoma. JAMA 307 (10): 1062-71, 2012.
  40. Molenaar JJ, Koster J, Zwijnenburg DA, et al.: Sequencing of neuroblastoma identifies chromothripsis and defects in neuritogenesis genes. Nature 483 (7391): 589-93, 2012.
  41. Sausen M, Leary RJ, Jones S, et al.: Integrated genomic analyses identify ARID1A and ARID1B alterations in the childhood cancer neuroblastoma. Nat Genet 45 (1): 12-7, 2013.
  42. Bresler SC, Weiser DA, Huwe PJ, et al.: ALK mutations confer differential oncogenic activation and sensitivity to ALK inhibition therapy in neuroblastoma. Cancer Cell 26 (5): 682-94, 2014.
  43. Janoueix-Lerosey I, Lequin D, Brugières L, et al.: Somatic and germline activating mutations of the ALK kinase receptor in neuroblastoma. Nature 455 (7215): 967-70, 2008.
  44. Eleveld TF, Oldridge DA, Bernard V, et al.: Relapsed neuroblastomas show frequent RAS-MAPK pathway mutations. Nat Genet 47 (8): 864-71, 2015.
  45. Schramm A, Köster J, Assenov Y, et al.: Mutational dynamics between primary and relapse neuroblastomas. Nat Genet 47 (8): 872-7, 2015.
  46. Bellini A, Bernard V, Leroy Q, et al.: Deep Sequencing Reveals Occurrence of Subclonal ALK Mutations in Neuroblastoma at Diagnosis. Clin Cancer Res 21 (21): 4913-21, 2015.
  47. Kurihara S, Hiyama E, Onitake Y, et al.: Clinical features of ATRX or DAXX mutated neuroblastoma. J Pediatr Surg 49 (12): 1835-8, 2014.
  48. Mac SM, D'Cunha CA, Farnham PJ: Direct recruitment of N-myc to target gene promoters. Mol Carcinog 29 (2): 76-86, 2000.
  49. Wang LL, Teshiba R, Ikegaki N, et al.: Augmented expression of MYC and/or MYCN protein defines highly aggressive MYC-driven neuroblastoma: a Children's Oncology Group study. Br J Cancer 113 (1): 57-63, 2015.
  50. Suganuma R, Wang LL, Sano H, et al.: Peripheral neuroblastic tumors with genotype-phenotype discordance: a report from the Children's Oncology Group and the International Neuroblastoma Pathology Committee. Pediatr Blood Cancer 60 (3): 363-70, 2013.
  51. Maris JM, Matthay KK: Molecular biology of neuroblastoma. J Clin Oncol 17 (7): 2264-79, 1999.
  52. Forlenza CJ, Boudreau JE, Zheng J, et al.: KIR3DL1 Allelic Polymorphism and HLA-B Epitopes Modulate Response to Anti-GD2 Monoclonal Antibody in Patients With Neuroblastoma. J Clin Oncol 34 (21): 2443-51, 2016.
  53. Venstrom JM, Zheng J, Noor N, et al.: KIR and HLA genotypes are associated with disease progression and survival following autologous hematopoietic stem cell transplantation for high-risk neuroblastoma. Clin Cancer Res 15 (23): 7330-4, 2009.
  54. Citak C, Karadeniz C, Dalgic B, et al.: Intestinal lymphangiectasia as a first manifestation of neuroblastoma. Pediatr Blood Cancer 46 (1): 105-7, 2006.
  55. Bourdeaut F, de Carli E, Timsit S, et al.: VIP hypersecretion as primary or secondary syndrome in neuroblastoma: A retrospective study by the Société Française des Cancers de l'Enfant (SFCE). Pediatr Blood Cancer 52 (5): 585-90, 2009.
  56. Mahoney NR, Liu GT, Menacker SJ, et al.: Pediatric horner syndrome: etiologies and roles of imaging and urine studies to detect neuroblastoma and other responsible mass lesions. Am J Ophthalmol 142 (4): 651-9, 2006.
  57. Conte M, Parodi S, De Bernardi B, et al.: Neuroblastoma in adolescents: the Italian experience. Cancer 106 (6): 1409-17, 2006.
  58. Matthay KK, Blaes F, Hero B, et al.: Opsoclonus myoclonus syndrome in neuroblastoma a report from a workshop on the dancing eyes syndrome at the advances in neuroblastoma meeting in Genoa, Italy, 2004. Cancer Lett 228 (1-2): 275-82, 2005.
  59. Rudnick E, Khakoo Y, Antunes NL, et al.: Opsoclonus-myoclonus-ataxia syndrome in neuroblastoma: clinical outcome and antineuronal antibodies-a report from the Children's Cancer Group Study. Med Pediatr Oncol 36 (6): 612-22, 2001.
  60. Pranzatelli MR: The neurobiology of the opsoclonus-myoclonus syndrome. Clin Neuropharmacol 15 (3): 186-228, 1992.
  61. Mitchell WG, Davalos-Gonzalez Y, Brumm VL, et al.: Opsoclonus-ataxia caused by childhood neuroblastoma: developmental and neurologic sequelae. Pediatrics 109 (1): 86-98, 2002.
  62. Connolly AM, Pestronk A, Mehta S, et al.: Serum autoantibodies in childhood opsoclonus-myoclonus syndrome: an analysis of antigenic targets in neural tissues. J Pediatr 130 (6): 878-84, 1997.
  63. Cooper R, Khakoo Y, Matthay KK, et al.: Opsoclonus-myoclonus-ataxia syndrome in neuroblastoma: histopathologic features-a report from the Children's Cancer Group. Med Pediatr Oncol 36 (6): 623-9, 2001.
  64. Russo C, Cohn SL, Petruzzi MJ, et al.: Long-term neurologic outcome in children with opsoclonus-myoclonus associated with neuroblastoma: a report from the Pediatric Oncology Group. Med Pediatr Oncol 28 (4): 284-8, 1997.
  65. Bell J, Moran C, Blatt J: Response to rituximab in a child with neuroblastoma and opsoclonus-myoclonus. Pediatr Blood Cancer 50 (2): 370-1, 2008.
  66. Corapcioglu F, Mutlu H, Kara B, et al.: Response to rituximab and prednisolone for opsoclonus-myoclonus-ataxia syndrome in a child with ganglioneuroblastoma. Pediatr Hematol Oncol 25 (8): 756-61, 2008.
  67. Vik TA, Pfluger T, Kadota R, et al.: (123)I-mIBG scintigraphy in patients with known or suspected neuroblastoma: Results from a prospective multicenter trial. Pediatr Blood Cancer 52 (7): 784-90, 2009.
  68. Yang J, Codreanu I, Servaes S, et al.: I-131 MIBG post-therapy scan is more sensitive than I-123 MIBG pretherapy scan in the evaluation of metastatic neuroblastoma. Nucl Med Commun 33 (11): 1134-7, 2012.
  69. Jennings RW, LaQuaglia MP, Leong K, et al.: Fetal neuroblastoma: prenatal diagnosis and natural history. J Pediatr Surg 28 (9): 1168-74, 1993.
  70. Nuchtern JG, London WB, Barnewolt CE, et al.: A prospective study of expectant observation as primary therapy for neuroblastoma in young infants: a Children's Oncology Group study. Ann Surg 256 (4): 573-80, 2012.
  71. Hero B, Simon T, Spitz R, et al.: Localized infant neuroblastomas often show spontaneous regression: results of the prospective trials NB95-S and NB97. J Clin Oncol 26 (9): 1504-10, 2008.
  72. Brodeur GM, Pritchard J, Berthold F, et al.: Revisions of the international criteria for neuroblastoma diagnosis, staging, and response to treatment. J Clin Oncol 11 (8): 1466-77, 1993.
  73. Horner MJ, Ries LA, Krapcho M, et al.: SEER Cancer Statistics Review, 1975-2006. Bethesda, Md: National Cancer Institute, 2009. Also available online. Last accessed April 04, 2017.
  74. Pinto NR, Applebaum MA, Volchenboum SL, et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma. J Clin Oncol 33 (27): 3008-17, 2015.
  75. Adams GA, Shochat SJ, Smith EI, et al.: Thoracic neuroblastoma: a Pediatric Oncology Group study. J Pediatr Surg 28 (3): 372-7; discussion 377-8, 1993.
  76. Evans AE, Albo V, D'Angio GJ, et al.: Factors influencing survival of children with nonmetastatic neuroblastoma. Cancer 38 (2): 661-6, 1976.
  77. Hayes FA, Green A, Hustu HO, et al.: Surgicopathologic staging of neuroblastoma: prognostic significance of regional lymph node metastases. J Pediatr 102 (1): 59-62, 1983.
  78. Cotterill SJ, Pearson AD, Pritchard J, et al.: Clinical prognostic factors in 1277 patients with neuroblastoma: results of The European Neuroblastoma Study Group 'Survey' 1982-1992. Eur J Cancer 36 (7): 901-8, 2000.
  79. Gustafson WC, Matthay KK: Progress towards personalized therapeutics: biologic- and risk-directed therapy for neuroblastoma. Expert Rev Neurother 11 (10): 1411-23, 2011.
  80. Isaacs H Jr: Fetal and neonatal neuroblastoma: retrospective review of 271 cases. Fetal Pediatr Pathol 26 (4): 177-84, 2007 Jul-Aug.
  81. Strother DR, London WB, Schmidt ML, et al.: Outcome after surgery alone or with restricted use of chemotherapy for patients with low-risk neuroblastoma: results of Children's Oncology Group study P9641. J Clin Oncol 30 (15): 1842-8, 2012.
  82. Baker DL, Schmidt ML, Cohn SL, et al.: Outcome after reduced chemotherapy for intermediate-risk neuroblastoma. N Engl J Med 363 (14): 1313-23, 2010.
  83. Castleberry RP, Kun LE, Shuster JJ, et al.: Radiotherapy improves the outlook for patients older than 1 year with Pediatric Oncology Group stage C neuroblastoma. J Clin Oncol 9 (5): 789-95, 1991.
  84. Bowman LC, Castleberry RP, Cantor A, et al.: Genetic staging of unresectable or metastatic neuroblastoma in infants: a Pediatric Oncology Group study. J Natl Cancer Inst 89 (5): 373-80, 1997.
  85. Castleberry RP, Shuster JJ, Altshuler G, et al.: Infants with neuroblastoma and regional lymph node metastases have a favorable outlook after limited postoperative chemotherapy: a Pediatric Oncology Group study. J Clin Oncol 10 (8): 1299-304, 1992.
  86. West DC, Shamberger RC, Macklis RM, et al.: Stage III neuroblastoma over 1 year of age at diagnosis: improved survival with intensive multimodality therapy including multiple alkylating agents. J Clin Oncol 11 (1): 84-90, 1993.
  87. Paul SR, Tarbell NJ, Korf B, et al.: Stage IV neuroblastoma in infants. Long-term survival. Cancer 67 (6): 1493-7, 1991.
  88. Bowman LC, Hancock ML, Santana VM, et al.: Impact of intensified therapy on clinical outcome in infants and children with neuroblastoma: the St Jude Children's Research Hospital experience, 1962 to 1988. J Clin Oncol 9 (9): 1599-608, 1991.
  89. Look AT, Hayes FA, Shuster JJ, et al.: Clinical relevance of tumor cell ploidy and N-myc gene amplification in childhood neuroblastoma: a Pediatric Oncology Group study. J Clin Oncol 9 (4): 581-91, 1991.
  90. Schmidt ML, Lukens JN, Seeger RC, et al.: Biologic factors determine prognosis in infants with stage IV neuroblastoma: A prospective Children's Cancer Group study. J Clin Oncol 18 (6): 1260-8, 2000.
  91. Schmidt ML, Lal A, Seeger RC, et al.: Favorable prognosis for patients 12 to 18 months of age with stage 4 nonamplified MYCN neuroblastoma: a Children's Cancer Group Study. J Clin Oncol 23 (27): 6474-80, 2005.
  92. George RE, London WB, Cohn SL, et al.: Hyperdiploidy plus nonamplified MYCN confers a favorable prognosis in children 12 to 18 months old with disseminated neuroblastoma: a Pediatric Oncology Group study. J Clin Oncol 23 (27): 6466-73, 2005.
  93. Mazzocco K, Defferrari R, Sementa AR, et al.: Genetic abnormalities in adolescents and young adults with neuroblastoma: A report from the Italian Neuroblastoma group. Pediatr Blood Cancer 62 (10): 1725-32, 2015.
  94. Mossé YP, Deyell RJ, Berthold F, et al.: Neuroblastoma in older children, adolescents and young adults: a report from the International Neuroblastoma Risk Group project. Pediatr Blood Cancer 61 (4): 627-35, 2014.
  95. Kushner BH, Kramer K, LaQuaglia MP, et al.: Neuroblastoma in adolescents and adults: the Memorial Sloan-Kettering experience. Med Pediatr Oncol 41 (6): 508-15, 2003.
  96. Franks LM, Bollen A, Seeger RC, et al.: Neuroblastoma in adults and adolescents: an indolent course with poor survival. Cancer 79 (10): 2028-35, 1997.
  97. Vo KT, Matthay KK, Neuhaus J, et al.: Clinical, biologic, and prognostic differences on the basis of primary tumor site in neuroblastoma: a report from the international neuroblastoma risk group project. J Clin Oncol 32 (28): 3169-76, 2014.
  98. Hiyama E, Yokoyama T, Hiyama K, et al.: Multifocal neuroblastoma: biologic behavior and surgical aspects. Cancer 88 (8): 1955-63, 2000.
  99. Kubota M, Suita S, Tajiri T, et al.: Analysis of the prognostic factors relating to better clinical outcome in ganglioneuroblastoma. J Pediatr Surg 35 (1): 92-5, 2000.
  100. Peuchmaur M, d'Amore ES, Joshi VV, et al.: Revision of the International Neuroblastoma Pathology Classification: confirmation of favorable and unfavorable prognostic subsets in ganglioneuroblastoma, nodular. Cancer 98 (10): 2274-81, 2003.
  101. Ikeda H, Iehara T, Tsuchida Y, et al.: Experience with International Neuroblastoma Staging System and Pathology Classification. Br J Cancer 86 (7): 1110-6, 2002.
  102. Teshiba R, Kawano S, Wang LL, et al.: Age-dependent prognostic effect by Mitosis-Karyorrhexis Index in neuroblastoma: a report from the Children's Oncology Group. Pediatr Dev Pathol 17 (6): 441-9, 2014 Nov-Dec.
  103. Burchill SA, Lewis IJ, Abrams KR, et al.: Circulating neuroblastoma cells detected by reverse transcriptase polymerase chain reaction for tyrosine hydroxylase mRNA are an independent poor prognostic indicator in stage 4 neuroblastoma in children over 1 year. J Clin Oncol 19 (6): 1795-801, 2001.
  104. Seeger RC, Reynolds CP, Gallego R, et al.: Quantitative tumor cell content of bone marrow and blood as a predictor of outcome in stage IV neuroblastoma: a Children's Cancer Group Study. J Clin Oncol 18 (24): 4067-76, 2000.
  105. Bochennek K, Esser R, Lehrnbecher T, et al.: Impact of minimal residual disease detection prior to autologous stem cell transplantation for post-transplant outcome in high risk neuroblastoma. Klin Padiatr 224 (3): 139-42, 2012.
  106. Yanik GA, Parisi MT, Shulkin BL, et al.: Semiquantitative mIBG scoring as a prognostic indicator in patients with stage 4 neuroblastoma: a report from the Children's oncology group. J Nucl Med 54 (4): 541-8, 2013.
  107. George RE, Perez-Atayde AR, Yao X, et al.: Tumor histology during induction therapy in patients with high-risk neuroblastoma. Pediatr Blood Cancer 59 (3): 506-10, 2012.
  108. Bagatell R, McHugh K, Naranjo A, et al.: Assessment of Primary Site Response in Children With High-Risk Neuroblastoma: An International Multicenter Study. J Clin Oncol 34 (7): 740-6, 2016.
  109. Nickerson HJ, Matthay KK, Seeger RC, et al.: Favorable biology and outcome of stage IV-S neuroblastoma with supportive care or minimal therapy: a Children's Cancer Group study. J Clin Oncol 18 (3): 477-86, 2000.
  110. Ambros PF, Brodeur GM: Concept of tumorigenesis and regression. In: Brodeur GM, Sawada T, Tsuchida Y: Neuroblastoma. New York, NY: Elsevier Science, 2000, pp 21-32.
  111. Hiyama E, Hiyama K, Yokoyama T, et al.: Correlating telomerase activity levels with human neuroblastoma outcomes. Nat Med 1 (3): 249-55, 1995.
  112. Hiyama E, Reynolds CP: Telomerase as a biological and prognostic marker in neuroblastoma. In: Brodeur GM, Sawada T, Tsuchida Y: Neuroblastoma. New York, NY: Elsevier Science, 2000, pp 159-174.
  113. Kitanaka C, Kato K, Ijiri R, et al.: Increased Ras expression and caspase-independent neuroblastoma cell death: possible mechanism of spontaneous neuroblastoma regression. J Natl Cancer Inst 94 (5): 358-68, 2002.
  114. Brodeur GM, Minturn JE, Ho R, et al.: Trk receptor expression and inhibition in neuroblastomas. Clin Cancer Res 15 (10): 3244-50, 2009.
  115. Yamamoto K, Ohta S, Ito E, et al.: Marginal decrease in mortality and marked increase in incidence as a result of neuroblastoma screening at 6 months of age: cohort study in seven prefectures in Japan. J Clin Oncol 20 (5): 1209-14, 2002.
  116. Okazaki T, Kohno S, Mimaya J, et al.: Neuroblastoma detected by mass screening: the Tumor Board's role in its treatment. Pediatr Surg Int 20 (1): 27-32, 2004.
  117. Fritsch P, Kerbl R, Lackner H, et al.: "Wait and see" strategy in localized neuroblastoma in infants: an option not only for cases detected by mass screening. Pediatr Blood Cancer 43 (6): 679-82, 2004.

Cellular Classification of Neuroblastic Tumors

Neuroblastomas are classified as one of the small round blue cell tumors of childhood. They are a heterogenous group of tumors composed of cellular aggregates with different degrees of differentiation, from mature ganglioneuromas to less mature ganglioneuroblastomas to immature neuroblastomas, reflecting the varying malignant potential of these tumors.[1]

There are two cellular classification systems for neuroblastoma:

  • International Neuroblastoma Pathology Classification (INPC) System.
  • International Neuroblastoma Risk Group (INRG) Classification System.

International Neuroblastoma Pathology Classification (INPC) System

The INPC system involves evaluation of tumor specimens obtained before therapy for the following morphologic features:[2,3,4,5,6]

  • Amount of Schwannian stroma.
  • Degree of neuroblastic maturation.
  • Mitosis-karyorrhexis index of the neuroblastic cells.

Favorable and unfavorable prognoses are defined on the basis of these histologic parameters and patient age. The prognostic significance of this classification system, and of related systems using similar criteria, has been confirmed in several studies.[2,3,4,6]

In the future, the INPC system is likely to be replaced by a system that does not include patient age as a part of cellular classification.

Table 1. Prognostic Evaluation of Neuroblastic Tumors According to the International Neuroblastoma Pathology Classification (Shimada System)a
International Neuroblastoma Pathology classificationOriginal Shimada classificationPrognostic group
MKI: mitosis-karyorrhexis index.
a Reprinted with permission. Copyright © 1999 American Cancer Society. All rights reserved.[2]Hiroyuki Shimada, Inge M. Ambros, Louis P. Dehner, Jun-ichi Hata, Vijay V. Joshi, Borghild Roald, Daniel O. Stram, Robert B. Gerbing, John N. Lukens, Katherine K. Matthay, Robert P. Castleberry, The International Neuroblastoma Pathology Classification (the Shimada System), Cancer, volume 86, issue 2, pages 364-72.
b Subtypes of neuroblastoma were described in detail elsewhere.[7]
c Rare subtype, especially diagnosed in this age group. Further investigation and analysis required.
d Prognostic grouping for these tumor categories is not related to patient age.
Neuroblastoma(Schwannian stroma-poor)bStroma-poor 
 FavorableFavorableFavorable
 <1.5 yrsPoorly differentiated or differentiating & low or intermediate MKI tumor  
 1.5-5 yrsDifferentiating & low MKI tumor  
 UnfavorableUnfavorableUnfavorable
 <1.5 yrsa) undifferentiated tumorc  
b) high MKI tumor
 1.5-5 yrsa) undifferentiated or poorly differentiated tumor  
b) intermediate or high MKI tumor
 ≥5 yrsAll tumors  
Ganglioneuroblastoma, intermixed(Schwannian stroma-rich)Stroma-rich Intermixed (favorable)Favorabled
Ganglioneuroma(Schwannian stroma-dominant)  
 Maturing Well differentiated (favorable)Favorabled
 Mature Ganglioneuroma 
Ganglioneuroblastoma, nodular(composite Schwannian stroma-rich/stroma-dominate and stroma-poor)Stroma-rich nodular (unfavorable)Unfavorabled

Most neuroblastomas with MYCN amplification in the INPC system also have unfavorable histology, but about 7% have favorable histology. Of those with MYCN amplification and favorable histology, most do not express MYCN, despite the gene being amplified, and have a more favorable prognosis than do those that do express MYCN.[8]

International Neuroblastoma Risk Group (INRG) Classification System

The INRG used a survival-tree analysis to compare 35 prognostic factors in more than 8,800 patients with neuroblastoma from a variety of clinical trials. The following INPC (Shimada system) histologic factors were included in the analysis:[9,10]

  • Diagnostic category.
  • Grade of differentiation.
  • Mitosis/karyorrhexis index.

Because patient age is used in all risk stratification systems, a cellular classification system that did not employ patient age was desirable, and underlying histologic criteria, rather than INPC or Shimada Classification, was used in the final decision tree. Histologic findings discriminated prognostic groups most clearly in two subsets of patients, as shown in Table 2.

Table 2. Histologic Discrimination of International Neuroblastoma Risk Group Subsets of Neuroblastoma Patientsa
INSS Stage/Histologic SubtypeNumber of CasesEFS (%)OS (%)
EFS = event-free survival; GN = ganglioneuroma; GNB = ganglioneuroblastoma; INSS = International Neuroblastoma Staging System; NB = neuroblastoma; OS = overall survival.
a Adapted from Cohn et al.[9]
INSS stage 1, 2, 3, 4S5,13183 ± 191 ± 1
 GN, maturing16297 ± 298 ± 2
GNB, intermixed
NB4,97083 ± 190 ± 1
GNB, nodular
INSS stage 2, 3; age >547 d26069 ± 381 ± 2
 11q normal and differentiating1680 ± 16100
11q aberration or undifferentiated4961 ± 1173 ± 11

The INRG histologic subsets are incorporated into the INRG Risk Classification Schema. (Refer to Table 6 in the Treatment Option Overview for Neuroblastoma section of this summary for more information.)

References:

  1. Joshi VV, Silverman JF: Pathology of neuroblastic tumors. Semin Diagn Pathol 11 (2): 107-17, 1994.
  2. Shimada H, Ambros IM, Dehner LP, et al.: The International Neuroblastoma Pathology Classification (the Shimada system). Cancer 86 (2): 364-72, 1999.
  3. Shimada H, Umehara S, Monobe Y, et al.: International neuroblastoma pathology classification for prognostic evaluation of patients with peripheral neuroblastic tumors: a report from the Children's Cancer Group. Cancer 92 (9): 2451-61, 2001.
  4. Goto S, Umehara S, Gerbing RB, et al.: Histopathology (International Neuroblastoma Pathology Classification) and MYCN status in patients with peripheral neuroblastic tumors: a report from the Children's Cancer Group. Cancer 92 (10): 2699-708, 2001.
  5. Peuchmaur M, d'Amore ES, Joshi VV, et al.: Revision of the International Neuroblastoma Pathology Classification: confirmation of favorable and unfavorable prognostic subsets in ganglioneuroblastoma, nodular. Cancer 98 (10): 2274-81, 2003.
  6. Teshiba R, Kawano S, Wang LL, et al.: Age-dependent prognostic effect by Mitosis-Karyorrhexis Index in neuroblastoma: a report from the Children's Oncology Group. Pediatr Dev Pathol 17 (6): 441-9, 2014 Nov-Dec.
  7. Shimada H, Ambros IM, Dehner LP, et al.: Terminology and morphologic criteria of neuroblastic tumors: recommendations by the International Neuroblastoma Pathology Committee. Cancer 86 (2): 349-63, 1999.
  8. Suganuma R, Wang LL, Sano H, et al.: Peripheral neuroblastic tumors with genotype-phenotype discordance: a report from the Children's Oncology Group and the International Neuroblastoma Pathology Committee. Pediatr Blood Cancer 60 (3): 363-70, 2013.
  9. Cohn SL, Pearson AD, London WB, et al.: The International Neuroblastoma Risk Group (INRG) classification system: an INRG Task Force report. J Clin Oncol 27 (2): 289-97, 2009.
  10. Okamatsu C, London WB, Naranjo A, et al.: Clinicopathological characteristics of ganglioneuroma and ganglioneuroblastoma: a report from the CCG and COG. Pediatr Blood Cancer 53 (4): 563-9, 2009.

Stage Information for Neuroblastoma

Staging Evaluation

Approximately 70% of patients with neuroblastoma have metastatic disease at diagnosis. A thorough evaluation for metastatic disease is performed before therapy initiation. The studies described below are typically performed.[1]

Metaiodobenzylguanidine (mIBG) scan

The extent of metastatic disease is assessed by mIBG scan, which is applicable to all sites of disease, including soft tissue, bone marrow, and cortical bone. Approximately 90% of neuroblastomas will be mIBG avid. The mIBG scan has a sensitivity and specificity of 90% to 99%, and mIBG avidity is equally distributed between primary and metastatic sites.[2] Although iodine 123 (123 I) has a shorter half-life, it is preferred over 131 I because of its lower radiation dose, better quality images, reduced thyroid toxicity, and lower cost. 18F-fluorodeoxyglucose positron emission tomography (PET) scans are used to evaluate extent of disease in patients with tumors that are not mIBG avid.[3]

Imaging with 123 I-mIBG is optimal for identifying soft tissue and bony metastases and was shown to be superior to PET-computed tomography (PET-CT) in one prospective comparison.[4] Baseline mIBG scans performed at diagnosis provide an excellent method for monitoring disease response and performing posttherapy surveillance.[5] A retrospective analysis of paired mIBG and PET scans in 60 newly diagnosed neuroblastoma patients demonstrated that for International Neuroblastoma Staging System (INSS) stages 1 and 2 patients, PET was superior at determining the extent of primary disease and more sensitive for detection of residual masses. In contrast, for stage 4 disease, 123 I-mIBG imaging was superior for the detection of bone marrow and bony metastases.[3]

Curie score and SIOPEN score

Multiple groups have investigated a semiquantitative scoring method to evaluate disease extent and prognostic value. The most common scoring methods in use for evaluation of disease extent and response are the Curie and the International Society of Paediatric Oncology Europe Neuroblastoma (SIOPEN) methods.

  • Curie score: The Curie score is a semiquantitative scoring system developed to predict the extent and severity of mIBG-avid disease. The use of the Curie scoring system was assessed as a prognostic marker for response and survival with mIBG-avid, stage 4, newly diagnosed, high-risk neuroblastoma (N = 280), treated on the Children's Oncology Group (COG) protocol COG-A3973 (NCT00004188). Patients with a Curie score higher than 2 after induction therapy had a significantly worse event-free survival (EFS) than did those with scores lower than 2 (3-year EFS, 15.4% ± 5.3% for Curie score >2 vs. 44.9% ± 3.9% for Curie score ≤2; P < .001). A postinduction Curie score higher than 2 identified a cohort of patients at higher risk of an event, independent of other known neuroblastoma factors, including age, MYCN status, ploidy, mitosis-karyorrhexis index, and histologic grade.[6]
  • SIOPEN score: SIOPEN independently developed an mIBG scan scoring system that divided the body into 12 segments, rather than 9 segments, and assigned six degrees, rather than four degrees, of mIBG uptake in each segment. A retrospective study of 58 stage 4 patients from the German Pediatric Oncology Group compared the prognostic value of the Curie and SIOPEN scoring methods. At diagnosis, a Curie score of 2 or lower and a SIOPEN score of 4 or lower (best cutoff) at diagnosis correlated to significantly better EFS and overall survival (OS), compared with higher scores. After four cycles of induction, those with complete response by mIBG had a better outcome than did those with residual uptake; however, after six cycles, there was no significant difference.[7]

Other staging tests and procedures

Other tests and procedures used to stage neuroblastoma include the following:

  • Bone marrow aspiration and biopsy: Bone marrow is assessed by bilateral iliac crest marrow aspirates and trephine (core) bone marrow biopsies to exclude bone marrow involvement. To be considered adequate, core biopsy specimens must contain at least 1 cm of marrow, excluding cartilage. Bone marrow sampling may not be necessary for tumors that are otherwise stage 1.[8]
  • Lymph node assessment: Palpable lymph nodes are clinically examined and histologically confirmed if indicated for staging.[1] CT, magnetic resonance imaging (MRI), or both are used to assess lymph nodes in regions that are not readily identified on physical exam.
  • CT and MRI scan:
    • Three-dimensional (3-D) imaging of the primary tumor and potential lymph node drainage sites is done using CT scans and/or MRI scans of the chest, abdomen, and pelvis. Ultrasound is generally considered suboptimal for accurate 3-D measurements.
    • Paraspinal tumors may extend through neural foramina to compress the spinal cord. Therefore, MRI of the spine adjacent to any paraspinal tumor is part of the staging evaluation.
    • A brain/orbit CT and/or MRI is performed if clinically indicated by examination and/or uptake on mIBG scan.

Lumbar puncture is avoided because central nervous system (CNS) metastasis at diagnosis is rare,[9] and lumbar puncture may be associated with an increased incidence of subsequent development of CNS metastasis.[10]

International Neuroblastoma Staging Systems

International Neuroblastoma Staging System (INSS)

The INSS combines certain features from each of the previously used Evans and Pediatric Oncology Group (POG) staging systems [1,11] and is described in Table 3. This represented the first step in harmonizing disease staging and risk stratification worldwide. The INSS is a surgical staging system that was developed in 1988 and used the extent of resection to stage patients. This led to some variability in stage assignments in different countries because of regional differences in surgical strategy and, potentially, because of limited access to experienced pediatric surgeons. As a result of further advances in the understanding of neuroblastoma biology and genetics, a risk classification system was developed that incorporates clinical and biological factors in addition to INSS stage to facilitate risk group and treatment assignment for COG studies.[1,11,12,13]

Table 3. The International Neuroblastoma Staging System (INSS)
Stage/Prognostic GroupDescription
mIBG = metaiodobenzylguanidine.
Stage 1Localized tumor with complete gross excision, with or without microscopic residual disease; representative ipsilateral lymph nodes negative for tumor microscopically (i.e., nodes attached to and removed with the primary tumor may be positive).
Stage 2ALocalized tumor with incomplete gross excision; representative ipsilateral nonadherent lymph nodes negative for tumor microscopically.
Stage 2BLocalized tumor with or without complete gross excision, with ipsilateral nonadherent lymph nodes positive for tumor. Enlarged contralateral lymph nodes must be negative microscopically
Stage 3Unresectable unilateral tumor infiltrating across the midline, with or without regional lymph node involvement; or localized unilateral tumor with contralateral regional lymph node involvement; or midline tumor with bilateral extension by infiltration (unresectable) or by lymph node involvement. The midline is defined as the vertebral column. Tumors originating on one side and crossing the midline must infiltrate to or beyond the opposite side of the vertebral column.
Stage 4Any primary tumor with dissemination to distant lymph nodes, bone, bone marrow, liver, skin, and/or other organs, except as defined for stage 4S.
Stage 4SLocalized primary tumor, as defined for stage 1, 2A, or 2B, with dissemination limited to skin, liver, and/or bone marrow (by definition limited to infants younger than 12 months).[14]Marrow involvement should be minimal (i.e., <10% of total nucleated cells identified as malignant by bone biopsy or by bone marrow aspirate). More extensive bone marrow involvement would be considered stage 4 disease. The results of the mIBG scan, if performed, should be negative for disease in the bone marrow.

The COG Neuroblastoma Risk Grouping that incorporates INSS is described in the Treatment Option Overview for Neuroblastoma section of this summary.

A study from the International Neuroblastoma Risk Group database found 146 patients with distant metastases limited to lymph nodes, termed stage 4N, who tended to have favorable-biology disease and a good outcome (5-year OS, 85%), which suggests that less-intensive therapy might be considered.[15]

International Neuroblastoma Risk Group Staging System (INRGSS)

The INRGSS is a preoperative staging system that was developed specifically for the INRG classification system. The extent of disease is determined by the presence or absence of image-defined risk factors (IDRFs) and/or metastatic tumor at the time of diagnosis, before any treatment or surgery. IDRFs are surgical risk factors, detected by imaging, which could potentially make total tumor excision risky or difficult at the time of diagnosis and increase the risk of surgical complications.

Table 4. International Neuroblastoma Risk Group Staging Systema
StageDescription
IDRFs = image-defined risk factors; INSS = International Neuroblastoma Staging System.
a Adapted from Monclair et al.[16];[17]
L1Localized tumor not involving vital structures as defined by the list of IDRFsa and confined to one body compartment.
L2Locoregional tumor with presence of one or more IDRFs.a
M Distant metastatic disease (except MS).
MSMetastatic disease in children younger than 18 months with metastases confined to skin, liver, and/or bone marrow. The primary tumor can be INSS stage 1, 2, or 3.

IDRFs include the following:[16]

  • Ipsilateral tumor extension within two body compartments: neck and chest; chest and abdomen; abdomen and pelvis.
  • Infiltration of adjacent organs/structures: pericardium, diaphragm, kidney, liver, duodeno-pancreatic block, mesentery.
  • Encasement of major vessels by tumor: vertebral artery, internal jugular vein, subclavian vessels, carotid artery, aorta, vena cava, major thoracic vessels, branches of the superior mesenteric artery at its root and the coeliac axis, iliac vessels.
  • Compression of trachea or central bronchi.
  • Encasement of brachial plexus.
  • Infiltration of porto-hepatic or hepato-duodenal ligament.
  • Infiltration of the costo-vertebral junction between T9 and T12.
  • Tumor crossing the sciatic notch.
  • Tumor invading renal pedicle.
  • Extension of tumor to base of skull.
  • Intraspinal tumor extension such that more than one-third of the spinal canal is invaded, leptomeningeal space is obliterated, or spinal cord MRI signal is abnormal.

The INRGSS has incorporated this staging system into a risk grouping system using multiple other parameters at diagnosis.[18] (Refer to Table 6 in the Treatment Option Overview for Neuroblastoma section of this summary for more information.)

The INRGSS simplifies stages into L1, L2, M, or MS (refer to Table 4 and the list of IDRFs for more information). Localized tumors are classified as stage L1 or L2 disease on the basis of whether 1 or more of the 20 IDRFs are present.[16] For example, in the case of spinal cord compression, an IDRF is present when more than one-third of the spinal canal in the axial plane is invaded, when the leptomeningeal spaces are not visible, or when the spinal cord magnetic resonance signal intensity is abnormal. The INRG collaboration has also defined techniques for detecting and quantifying neuroblastoma in bone marrow, both at diagnosis and after treatment. Quantification of this disease may result in more accurate assessment of response to treatment, but has not yet been applied to any clinical trials.[19]

By combining the INRGSS, preoperative imaging, and biological factors, each patient is assigned a risk stage that predicts outcome and dictates the appropriate treatment approach. The INRGSS has predictive value for children with INSS stage 1, 2, and 3, with stage L1 having a 5-year EFS of 90% and OS of 96%, versus 79% EFS and 89% OS for L2.[16] However, the INSS stage also discriminates among INRGSS stage L2 patients, with INSS stages 1, 2, and 3 (non-MYCN amplified) having 5-year EFS rates of 94%, 81%, and 76% and 5-year OS rates of 99%, 93%, and 83%, respectively. In the latter study, many children with L2 tumors underwent primary surgery and had an outcome significantly superior to that of children who underwent biopsy only as the initial operative procedure (5-year OS of 93% vs. 83%).[20] Many of the children entered on the latter study underwent primary surgery against protocol in spite of IDRFs and L2 classification, and these children had superior outcome (5-year OS of 95% vs. 83%). However, these children also had a 17% rate of operative complications (vs. 5%). In L1 patients undergoing primary surgery, those with operative complications had a lower OS (92% vs. 97%).[20]

Most international protocols have begun to incorporate the collection and use of IDRFs in risk stratification and assignment of therapy.[21,22] The COG has been collecting and evaluating INRGSS data since 2006. A COG trial that opened in 2014 uses the INRGSS along with input from the surgeon to determine therapy for patients with certain localized disease and for stage 4S patients. Note that the INSS allows patients up to age 12 months to be classified as stage 4S, while the INRGSS allows patients up to age 18 months to be staged as MS. The primary tumor in INSS stage 4S must be INSS stage 1 or 2, while the primary tumor in MS can be INSS stage 3. It is anticipated that the use of standardized nomenclature will contribute substantially to more uniform staging and thereby facilitate comparisons of clinical trials conducted in different parts of the world.

References:

  1. Brodeur GM, Pritchard J, Berthold F, et al.: Revisions of the international criteria for neuroblastoma diagnosis, staging, and response to treatment. J Clin Oncol 11 (8): 1466-77, 1993.
  2. Howman-Giles R, Shaw PJ, Uren RF, et al.: Neuroblastoma and other neuroendocrine tumors. Semin Nucl Med 37 (4): 286-302, 2007.
  3. Sharp SE, Shulkin BL, Gelfand MJ, et al.: 123I-MIBG scintigraphy and 18F-FDG PET in neuroblastoma. J Nucl Med 50 (8): 1237-43, 2009.
  4. Papathanasiou ND, Gaze MN, Sullivan K, et al.: 18F-FDG PET/CT and 123I-metaiodobenzylguanidine imaging in high-risk neuroblastoma: diagnostic comparison and survival analysis. J Nucl Med 52 (4): 519-25, 2011.
  5. Kushner BH, Kramer K, Modak S, et al.: Sensitivity of surveillance studies for detecting asymptomatic and unsuspected relapse of high-risk neuroblastoma. J Clin Oncol 27 (7): 1041-6, 2009.
  6. Yanik GA, Parisi MT, Shulkin BL, et al.: Semiquantitative mIBG scoring as a prognostic indicator in patients with stage 4 neuroblastoma: a report from the Children's oncology group. J Nucl Med 54 (4): 541-8, 2013.
  7. Decarolis B, Schneider C, Hero B, et al.: Iodine-123 metaiodobenzylguanidine scintigraphy scoring allows prediction of outcome in patients with stage 4 neuroblastoma: results of the Cologne interscore comparison study. J Clin Oncol 31 (7): 944-51, 2013.
  8. Russell HV, Golding LA, Suell MN, et al.: The role of bone marrow evaluation in the staging of patients with otherwise localized, low-risk neuroblastoma. Pediatr Blood Cancer 45 (7): 916-9, 2005.
  9. DuBois SG, Kalika Y, Lukens JN, et al.: Metastatic sites in stage IV and IVS neuroblastoma correlate with age, tumor biology, and survival. J Pediatr Hematol Oncol 21 (3): 181-9, 1999 May-Jun.
  10. Kramer K, Kushner B, Heller G, et al.: Neuroblastoma metastatic to the central nervous system. The Memorial Sloan-kettering Cancer Center Experience and A Literature Review. Cancer 91 (8): 1510-9, 2001.
  11. Brodeur GM, Seeger RC, Barrett A, et al.: International criteria for diagnosis, staging, and response to treatment in patients with neuroblastoma. J Clin Oncol 6 (12): 1874-81, 1988.
  12. Castleberry RP, Shuster JJ, Smith EI: The Pediatric Oncology Group experience with the international staging system criteria for neuroblastoma. Member Institutions of the Pediatric Oncology Group. J Clin Oncol 12 (11): 2378-81, 1994.
  13. Ikeda H, Iehara T, Tsuchida Y, et al.: Experience with International Neuroblastoma Staging System and Pathology Classification. Br J Cancer 86 (7): 1110-6, 2002.
  14. Taggart DR, London WB, Schmidt ML, et al.: Prognostic value of the stage 4S metastatic pattern and tumor biology in patients with metastatic neuroblastoma diagnosed between birth and 18 months of age. J Clin Oncol 29 (33): 4358-64, 2011.
  15. Morgenstern DA, London WB, Stephens D, et al.: Metastatic neuroblastoma confined to distant lymph nodes (stage 4N) predicts outcome in patients with stage 4 disease: A study from the International Neuroblastoma Risk Group Database. J Clin Oncol 32 (12): 1228-35, 2014.
  16. Monclair T, Brodeur GM, Ambros PF, et al.: The International Neuroblastoma Risk Group (INRG) staging system: an INRG Task Force report. J Clin Oncol 27 (2): 298-303, 2009.
  17. Brisse HJ, McCarville MB, Granata C, et al.: Guidelines for imaging and staging of neuroblastic tumors: consensus report from the International Neuroblastoma Risk Group Project. Radiology 261 (1): 243-57, 2011.
  18. Pinto NR, Applebaum MA, Volchenboum SL, et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma. J Clin Oncol 33 (27): 3008-17, 2015.
  19. Burchill SA, Beiske K, Shimada H, et al.: Recommendations for the standardization of bone marrow disease assessment and reporting in children with neuroblastoma on behalf of the International Neuroblastoma Response Criteria Bone Marrow Working Group. Cancer 123 (7): 1095-1105, 2017.
  20. Monclair T, Mosseri V, Cecchetto G, et al.: Influence of image-defined risk factors on the outcome of patients with localised neuroblastoma. A report from the LNESG1 study of the European International Society of Paediatric Oncology Neuroblastoma Group. Pediatr Blood Cancer 62 (9): 1536-42, 2015.
  21. Cecchetto G, Mosseri V, De Bernardi B, et al.: Surgical risk factors in primary surgery for localized neuroblastoma: the LNESG1 study of the European International Society of Pediatric Oncology Neuroblastoma Group. J Clin Oncol 23 (33): 8483-9, 2005.
  22. Simon T, Hero B, Benz-Bohm G, et al.: Review of image defined risk factors in localized neuroblastoma patients: Results of the GPOH NB97 trial. Pediatr Blood Cancer 50 (5): 965-9, 2008.

Treatment Option Overview for Neuroblastoma

Previously, most children with neuroblastoma in North America were treated according to the Children's Oncology Group (COG) risk-group assignment, even if they were not enrolled in a COG study. In the most recent COG study, the International Neuroblastoma Risk Group (INRG) system was used to assign treatment. Because the older system is still being used by some physicians to plan treatment, the treatments described in this summary are based on both the INRG system and the COG risk stratification system. In the INRG system, each child is assigned to a group according to the presence or absence of image-defined risk factors and metastasis. (Refer to the list of image-defined risk factors [IDRFs] in the Stage Information for Neuroblastoma section of this summary for more information.) Ongoing COG clinical trials have incorporated the International Neuroblastoma Risk Group Staging System (INRGSS) in lieu of the International Neuroblastoma Staging System (INSS). In the previous COG risk system, each child was assigned to a low-risk, intermediate-risk, or high-risk group (refer to Tables 7, 10, and 13 for more information) based on the following:[1,2,3,4,5,6]

  • INSS stage.
  • Age.
  • International Neuroblastoma Pathologic Classification (INPC).
  • Ploidy.
  • Amplification of the MYCN oncogene within tumor tissue.[1,2,3,4,5,6]

Other biological factors that influenced treatment selection in previous COG studies included unbalanced 11q loss of heterozygosity and loss of heterozygosity for chromosome 1p.[7,8] However, in 2012, the COG Neuroblastoma Committee defined favorable genomics, for purposes of risk assignment, as hyperdiploid neuroblastoma cells without segmental copy number aberrations, including no loss of copy number at 1p, 3p, 4p, or 11q and no gain of copy number at 1q, 2p, or 17q.

The treatment of neuroblastoma has evolved over the past 60 years. Generally, treatment is based on whether the tumor is low, intermediate, or high risk:

  • For patients with low-risk tumors, the approach is either observation or resection. Five-year overall survival (OS) was 97% in a large COG study.[9]
  • For patients with intermediate-risk tumors, chemotherapy is often given before definitive resection, with the amount and duration based on clinical and tumor biological risk factors and response to therapy. In recent studies, select patients have been observed without undergoing chemotherapy or attempted resection. The 3-year OS rate for intermediate-risk patients was about 96% in a large COG study;[10] thus, the current trend is to decrease chemotherapy to diminish side effects.
  • For high-risk patients, treatment has intensified to include chemotherapy, surgery, radiation therapy, myeloablative therapy and stem cell transplantation, isotretinoin, and immunotherapy, resulting in survival rates of 40% to 50%.
Table 5. Treatment Options for Neuroblastoma
COG Risk-Group AssignmentTreatment Options
COG = Children's Oncology Group; GM-CSF = granulocyte-macrophage colony-stimulating factor;131 I-mIBG = iodine 131-metaiodobenzylguanidine; SCT = stem cell transplant.
Low-Risk NeuroblastomaSurgery followed by observation.
Chemotherapy with or without surgery(for symptomatic disease or unresectable progressive disease after surgery).
Observation without biopsy(for perinatal neuroblastoma with small adrenal tumors).
Radiation therapy(only for emergency therapy).
Intermediate-Risk NeuroblastomaChemotherapy with or without surgery.
Surgery and observation(in infants).
Radiation therapy(only for emergency therapy).
High-Risk NeuroblastomaA regimen of chemotherapy, surgery, myeloablative therapy and SCT, radiation therapy, and dinutuximab, with interleukin-2/GM-CSF and isotretinoin.
Stage 4S NeuroblastomaObservation with supportive care(for asymptomatic patients with favorable tumor biology).
Chemotherapy(for symptomatic patients, very young infants, or those with unfavorable biology).
Recurrent NeuroblastomaLocoregional recurrence in patients initially classified as low riskSurgery followed by observation or chemotherapy.
Chemotherapy that may be followed by surgery.
Metastatic recurrence in patients initially classified as low riskObservation(if metastatic disease is in a 4S pattern in an infant).
Chemotherapy.
Surgery followed by chemotherapy.
High-risk therapy.
Locoregional recurrence in patients initially classified as intermediate riskSurgery(complete resection).
Surgery (incomplete resection) followed by chemotherapy.
Metastatic recurrence in patients initially classified as intermediate riskHigh-risk therapy.
Recurrence in patients initially classified as high riskChemotherapy with or without immunotherapy.
131I-mIBG alone, in combination with other therapy, or followed by stem cell rescue.
Second autologous SCT after additional chemotherapy.
Novel therapeutic approaches
Recurrence in the central nervous systemSurgery and radiation therapy.
Novel therapeutic approaches.

Children's Oncology Group (COG) Neuroblastoma Risk Grouping

The treatment section of this document is organized to correspond with the COG risk-based treatment plan that assigned all patients to a low-, intermediate-, or high-risk group. The COG risk-based treatment plan is no longer in use, as current studies are based on the INRG risk grouping. This risk-based schema was based on the following factors:

  • Patient age at diagnosis.
  • Certain biological characteristics of the tumor, which included MYCN status, INPC histopathology classification, and tumor DNA index.
  • Stage of the tumor as defined by the INSS.

Table 7 (in the Treatment of Low-Risk Neuroblastoma section), Table 10 (in the Treatment of Intermediate-Risk Neuroblastoma section), and Table 13 (in the Treatment of High-Risk Neuroblastoma section) describe the risk-group assignment criteria used to assign treatment in the COG-P9641, COG-A3961, and COG-A3973 studies, respectively.

Assessment of risk for low-stage MYCN-amplified neuroblastoma is controversial because it is so rare. A study of 87 INSS stage 1 and 2 patients pooled from several clinical trial groups demonstrated no effect of age, stage, or initial treatment on outcome. The event-free survival (EFS) rate was 53% and the OS rate was 72%. Survival was superior in patients whose tumors were hyperdiploid, rather than diploid (EFS, 82% ± 20% vs. 37% ± 21%; OS, 94% ± 11% vs. 54% ± 15%).[11] The overall EFS and OS for infants with stage 4 and 4S disease and MYCN-amplification was only 30% at 2 to 5 years after treatment in a European study.[12] The COG considers infants with stage 4 and stage 4S disease with MYCN amplification to be at high risk.[4]

International Neuroblastoma Risk Grouping

The INRG classifies all neuroblastoma patients into 16 pretreatment risk groups on the basis of INRG stage, age, histologic category, grade of tumor differentiation, MYCN amplification, 11q aberration (a single segmental chromosomal aberration), and ploidy. They assigned four levels of risk according to outcomes among 8,800 patients with high-quality data, as they had been entered on clinical trials (refer to Table 6 below). In the overall risk grouping, histology is an important risk determinant for all stage L1 and L2 tumors, and grade of differentiation discriminates among neuroblastomas and nodular ganglioneuroblastomas in patients older than 18 months. The goals are to develop shared data from the patients and defined risk groups for future trials.[13]

Table 6. International Neuroblastoma Risk Group (INRG) Pretreatment Classification Schemaa
INRG StageHistologic CategoryGrade of Tumor DifferentiationMYCN11q AberrationPloidyPretreatment Risk Group
GN = ganglioneuroma; GNB = ganglioneuroblastoma; NA = not amplified.
a Reprinted with permission. © (2015) American Society of Clinical Oncology. All rights reserved. Pinto N et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma, J Clin Oncol 33 (27), 2015: 3008-3017.[14]
L1/L2GN maturing, GNB intermixed    A (very low)
L1Any, except GN maturing or GNB intermixed NA  B (very low)
Amplified  K (high)
L2 
 Age <18 moAny, except GN maturing or GNB intermixed NANo D (low)
Yes G (intermediate)
 Age ≥18 moGNB nodular neuroblastomaDifferentiatingNANo E (low)
Yes H (intermediate)
Poorly differentiated or undifferentiatedNA  H (intermediate)
Amplified  N (high)
M 
 Age <18 mo  NA HyperdiploidF (low)
 Age <12 mo  NA DiploidI (intermediate)
 Age 12 to <18 mo  NA DiploidJ (intermediate)
 Age <18 mo  Amplified  O (high)
 Age ≥18 mo     P (high)
MS 
 Age <18 mo  NANo C (very low)
Yes Q (high)
Amplified  R (high)

Controversy exists regarding the previous COG risk grouping system, the INRG Risk Grouping Schema in current use, and the treatment of certain small subsets of patients.[15,16,17] Risk group assignment and recommended treatment are expected to evolve as additional outcome data are analyzed. For example, the risk group assignment for INSS stage 4 neuroblastoma in patients aged 12 to 18 months changed in 2005 for those whose tumors had single-copy MYCN and all favorable biological features; these patients had been previously classified as high risk, but data from both Pediatric Oncology Group and Children's Cancer Group studies suggested that this subgroup of patients could be successfully treated as intermediate risk.[18,19,20] Future versions of the INRG Risk Grouping are expected to contain more tumor genomic criteria to help assign risk.[14]

Description of International Neuroblastoma Response Criteria

Before therapy can be stopped after the initially planned number of cycles, certain response criteria, depending on risk group and treatment assignment, must be met. These criteria are defined as follows:[21,22]

  • Complete Response: Total disappearance of tumor, with no evidence of disease, including resolution of mIBG uptake. Vanillylmandelic acid (VMA) and homovanillic acid (HVA) are normal.
  • Very Good Partial Response: Primary tumor has decreased by 90% to 99%, and no evidence of metastatic disease, including resolution of mIBG uptake. Urine VMA/HVA are normal. Residual bone scan changes are allowed.
  • Partial Response: 50% to 90% decrease in the size of all measurable lesions; the number of mIBG scan-positive sites is decreased by more than 50% and no new lesions are present; no more than one positive bone marrow site allowed if this represents a reduction in the number of sites originally positive for tumor at diagnosis.
  • Mixed Response: No new lesions, 50% to 90% reduction of any measurable lesion (primary or metastatic) with less than 50% reduction in other lesions and less than 25% increase in any lesion.
  • No Response or Stable Disease: No new lesions; less than 50% reduction and less than 25% increase in any lesion.
  • Progressive Disease: Any new lesion; increase in any measurable lesion by more than 25%; previous negative bone marrow now positive for tumor.

    Persistent elevation in urinary VMA/HVA with stable disease or an increase in VMA/HVA without clinical or radiographic evidence of progression does not indicate progressive disease, but warrants continued follow-up.

    Care should be taken in interpreting the development of metastatic disease in an infant who was initially considered to have stage 1 or 2 disease. If the pattern of metastases in such a patient is consistent with a 4S pattern of disease (skin, liver, bone marrow less than 10% involved), these patients are not classified as having progressive/metastatic disease, which would typically be a criteria for removal from protocol therapy. Instead, these patients are managed as stage 4S.

Controversy exists regarding the necessity of measuring the primary tumor response in all three dimensions or whether the single longest dimension, as in Response Evaluation Criteria In Solid Tumors (RECIST) tumor response determination, is just as useful.[23]

Surgery

In patients without metastatic disease, the standard of care is to perform an initial surgery to accomplish the following:

  • Establish the diagnosis.
  • Resect as much of the primary tumor as is safely possible.
  • Accurately stage disease through sampling of regional lymph nodes that are not adherent to the tumor.
  • Obtain adequate tissue for biological studies.

In patients with L1 tumors (defined as having no image-defined surgical risk factors), resection is less likely to result in surgical complications and, generally, the tumors have been resected. L2 tumors, which have at least one image-defined surgical risk factor, have been treated with chemotherapy when deemed too risky to attempt resection, followed by surgery when the tumors have responded. Recent German studies of selected groups of patients have biopsied tissue and observed infants with both L1 and L2 tumors without MYCN amplification, avoiding additional surgery and chemotherapy in most patients.[24]

The COG reported that expectant observation in infants younger than 6 months with small (L1) adrenal masses resulted in an excellent EFS and OS while avoiding surgical intervention in a large majority of patients.[25] According to the surgical guidelines described in the intermediate-risk neuroblastoma clinical trial (ANBL0531 [NCT00499616]), the primary tumor is not routinely resected in patients with 4S neuroblastoma.

Whether there is any advantage to gross-total resection of the primary tumor mass after chemotherapy in stage 4 patients older than 18 months remains controversial.[26,27,28,29,30,31] A meta-analysis of stage 3 versus stage 4 neuroblastoma patients, at all ages combined, found an advantage for gross-total resection over subtotal resection in stage 3 neuroblastoma only, not stage 4.[32] Also, a small study suggested that after neoadjuvant chemotherapy, completeness of resection was affected by the number of IDRFs remaining.[33]

Radiation Therapy

In the completed COG treatment plan, radiation therapy for patients with low-risk or intermediate-risk neuroblastoma was reserved for symptomatic life-threatening or organ-threatening tumor bulk that did not respond rapidly enough to chemotherapy. Common situations in which radiation therapy is used in these patients include the following:

  • Infants aged 60 days and younger with stage 4S and marked respiratory compromise from liver metastases that has not responded to chemotherapy.
  • Symptomatic spinal cord compression that has not responded to initial chemotherapy and/or surgical decompression.

Treatment of Spinal Cord Compression

Spinal cord compression is considered a medical emergency. Immediate treatment is given because neurologic recovery is more likely when symptoms are present for a relatively short period of time before diagnosis and treatment. Recovery also depends on the severity of neurologic defects (weakness vs. paralysis). Neurologic outcome appears to be similar whether cord compression is treated with chemotherapy, radiation therapy, or surgery, although radiation therapy is used less frequently than in the past.

The completed COG low-risk and intermediate-risk neuroblastoma clinical trials recommended immediate chemotherapy for cord compression in low-risk or intermediate-risk patients.[34,35,36]

Children with severe spinal cord compression that does not promptly improve or those with worsening symptoms may benefit from neurosurgical intervention. Laminectomy may result in later kyphoscoliosis and may not eliminate the need for chemotherapy.[34,35,36] It was thought that osteoplastic laminotomy, a procedure that does not remove bone, would result in less spinal deformity. Osteoplastic laminotomy may be associated with a lower incidence of progressive spinal deformity requiring fusion, but there is no evidence that functional deficit is improved with laminoplasty.[37]

In a series of 34 infants with symptomatic epidural spinal cord compression, both surgery and chemotherapy provided unsatisfactory results once paraplegia had been established. The frequency of grade 3 motor deficits and bowel dysfunction increased with a longer symptom duration interval. Most infants with symptomatic epidural spinal cord compression developed sequelae, which were severe in about one-half of them.[38] This finding supports the need for greater awareness and timely intervention in these infants.

Surveillance During and After Treatment

Surveillance studies during and after treatment are able to detect asymptomatic and unsuspected relapse in a substantial portion of patients. In an overall surveillance plan, which includes urinary VMA/HVA testing, one of the most reliable imaging tests to detect disease progression or recurrence is the 123 I-metaiodobenzylguanidine scan.[39,40] Cross-sectional imaging with computed tomography scans is controversial because of the amount of radiation received and the low proportion of relapses detected with this modality.[41]

Special Considerations for the Treatment of Children With Cancer

Fortunately, cancer in children and adolescents is rare, although the overall incidence of childhood cancer has been slowly increasing since 1975.[42] Children and adolescents with cancer are usually referred to medical centers that have a multidisciplinary team of cancer specialists with experience treating the cancers that occur during childhood and adolescence. This multidisciplinary team approach incorporates the skills of the following health care professionals and others to ensure that children receive treatment, supportive care, and rehabilitation that will enable them to achieve optimal survival and quality of life:

  • Primary care physician.
  • Pediatric pathologists.
  • Pediatric surgeons.
  • Pediatric radiation oncologists.
  • Pediatric medical oncologists/hematologists.
  • Pediatric nurse specialists.
  • Social workers.
  • Child life professionals.

(Refer to the PDQ summaries on Supportive and Palliative Care for specific information about supportive care for children and adolescents with cancer.)

Guidelines for pediatric cancer centers and their role in the treatment of pediatric patients with cancer have been outlined by the American Academy of Pediatrics.[43] At these pediatric cancer centers, clinical trials are available for most types of cancer that occur in children and adolescents, and the opportunity to participate in these trials is offered to most patients and families. Clinical trials for children and adolescents with cancer are generally designed to compare potentially better therapy with therapy that is currently accepted as standard. Most of the progress made in identifying curative therapies for childhood cancers has been achieved through clinical trials. Other types of clinical trials explore or define novel therapies when there is no standard therapy for a cancer diagnosis. Information about ongoing clinical trials is available from the NCI website.

References:

  1. Cotterill SJ, Pearson AD, Pritchard J, et al.: Clinical prognostic factors in 1277 patients with neuroblastoma: results of The European Neuroblastoma Study Group 'Survey' 1982-1992. Eur J Cancer 36 (7): 901-8, 2000.
  2. Moroz V, Machin D, Faldum A, et al.: Changes over three decades in outcome and the prognostic influence of age-at-diagnosis in young patients with neuroblastoma: a report from the International Neuroblastoma Risk Group Project. Eur J Cancer 47 (4): 561-71, 2011.
  3. Look AT, Hayes FA, Shuster JJ, et al.: Clinical relevance of tumor cell ploidy and N-myc gene amplification in childhood neuroblastoma: a Pediatric Oncology Group study. J Clin Oncol 9 (4): 581-91, 1991.
  4. Schmidt ML, Lukens JN, Seeger RC, et al.: Biologic factors determine prognosis in infants with stage IV neuroblastoma: A prospective Children's Cancer Group study. J Clin Oncol 18 (6): 1260-8, 2000.
  5. Berthold F, Trechow R, Utsch S, et al.: Prognostic factors in metastatic neuroblastoma. A multivariate analysis of 182 cases. Am J Pediatr Hematol Oncol 14 (3): 207-15, 1992.
  6. Matthay KK, Perez C, Seeger RC, et al.: Successful treatment of stage III neuroblastoma based on prospective biologic staging: a Children's Cancer Group study. J Clin Oncol 16 (4): 1256-64, 1998.
  7. Attiyeh EF, London WB, Mossé YP, et al.: Chromosome 1p and 11q deletions and outcome in neuroblastoma. N Engl J Med 353 (21): 2243-53, 2005.
  8. Spitz R, Hero B, Simon T, et al.: Loss in chromosome 11q identifies tumors with increased risk for metastatic relapses in localized and 4S neuroblastoma. Clin Cancer Res 12 (11 Pt 1): 3368-73, 2006.
  9. Strother DR, London WB, Schmidt ML, et al.: Outcome after surgery alone or with restricted use of chemotherapy for patients with low-risk neuroblastoma: results of Children's Oncology Group study P9641. J Clin Oncol 30 (15): 1842-8, 2012.
  10. Baker DL, Schmidt ML, Cohn SL, et al.: Outcome after reduced chemotherapy for intermediate-risk neuroblastoma. N Engl J Med 363 (14): 1313-23, 2010.
  11. Bagatell R, Beck-Popovic M, London WB, et al.: Significance of MYCN amplification in international neuroblastoma staging system stage 1 and 2 neuroblastoma: a report from the International Neuroblastoma Risk Group database. J Clin Oncol 27 (3): 365-70, 2009.
  12. Canete A, Gerrard M, Rubie H, et al.: Poor survival for infants with MYCN-amplified metastatic neuroblastoma despite intensified treatment: the International Society of Paediatric Oncology European Neuroblastoma Experience. J Clin Oncol 27 (7): 1014-9, 2009.
  13. Cohn SL, Pearson AD, London WB, et al.: The International Neuroblastoma Risk Group (INRG) classification system: an INRG Task Force report. J Clin Oncol 27 (2): 289-97, 2009.
  14. Pinto NR, Applebaum MA, Volchenboum SL, et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma. J Clin Oncol 33 (27): 3008-17, 2015.
  15. Kushner BH, Cheung NK: Treatment reduction for neuroblastoma. Pediatr Blood Cancer 43 (6): 619-21, 2004.
  16. Kushner BH, Kramer K, LaQuaglia MP, et al.: Liver involvement in neuroblastoma: the Memorial Sloan-Kettering Experience supports treatment reduction in young patients. Pediatr Blood Cancer 46 (3): 278-84, 2006.
  17. Navarro S, Amann G, Beiske K, et al.: Prognostic value of International Neuroblastoma Pathology Classification in localized resectable peripheral neuroblastic tumors: a histopathologic study of localized neuroblastoma European Study Group 94.01 Trial and Protocol. J Clin Oncol 24 (4): 695-9, 2006.
  18. Schmidt ML, Lal A, Seeger RC, et al.: Favorable prognosis for patients 12 to 18 months of age with stage 4 nonamplified MYCN neuroblastoma: a Children's Cancer Group Study. J Clin Oncol 23 (27): 6474-80, 2005.
  19. London WB, Castleberry RP, Matthay KK, et al.: Evidence for an age cutoff greater than 365 days for neuroblastoma risk group stratification in the Children's Oncology Group. J Clin Oncol 23 (27): 6459-65, 2005.
  20. George RE, London WB, Cohn SL, et al.: Hyperdiploidy plus nonamplified MYCN confers a favorable prognosis in children 12 to 18 months old with disseminated neuroblastoma: a Pediatric Oncology Group study. J Clin Oncol 23 (27): 6466-73, 2005.
  21. Brodeur GM, Pritchard J, Berthold F, et al.: Revisions of the international criteria for neuroblastoma diagnosis, staging, and response to treatment. J Clin Oncol 11 (8): 1466-77, 1993.
  22. Brodeur GM, Seeger RC, Barrett A, et al.: International criteria for diagnosis, staging, and response to treatment in patients with neuroblastoma. J Clin Oncol 6 (12): 1874-81, 1988.
  23. Bagatell R, McHugh K, Naranjo A, et al.: Assessment of Primary Site Response in Children With High-Risk Neuroblastoma: An International Multicenter Study. J Clin Oncol 34 (7): 740-6, 2016.
  24. Hero B, Simon T, Spitz R, et al.: Localized infant neuroblastomas often show spontaneous regression: results of the prospective trials NB95-S and NB97. J Clin Oncol 26 (9): 1504-10, 2008.
  25. Nuchtern JG, London WB, Barnewolt CE, et al.: A prospective study of expectant observation as primary therapy for neuroblastoma in young infants: a Children's Oncology Group study. Ann Surg 256 (4): 573-80, 2012.
  26. Adkins ES, Sawin R, Gerbing RB, et al.: Efficacy of complete resection for high-risk neuroblastoma: a Children's Cancer Group study. J Pediatr Surg 39 (6): 931-6, 2004.
  27. Castel V, Tovar JA, Costa E, et al.: The role of surgery in stage IV neuroblastoma. J Pediatr Surg 37 (11): 1574-8, 2002.
  28. La Quaglia MP, Kushner BH, Su W, et al.: The impact of gross total resection on local control and survival in high-risk neuroblastoma. J Pediatr Surg 39 (3): 412-7; discussion 412-7, 2004.
  29. Simon T, Häberle B, Hero B, et al.: Role of surgery in the treatment of patients with stage 4 neuroblastoma age 18 months or older at diagnosis. J Clin Oncol 31 (6): 752-8, 2013.
  30. Englum BR, Rialon KL, Speicher PJ, et al.: Value of surgical resection in children with high-risk neuroblastoma. Pediatr Blood Cancer 62 (9): 1529-35, 2015.
  31. von Allmen D, Davidoff AM, London WB, et al.: Impact of Extent of Resection on Local Control and Survival in Patients From the COG A3973 Study With High-Risk Neuroblastoma. J Clin Oncol 35 (2): 208-216, 2017.
  32. Mullassery D, Farrelly P, Losty PD: Does aggressive surgical resection improve survival in advanced stage 3 and 4 neuroblastoma? A systematic review and meta-analysis. Pediatr Hematol Oncol 31 (8): 703-16, 2014.
  33. Irtan S, Brisse HJ, Minard-Colin V, et al.: Image-defined risk factor assessment of neurogenic tumors after neoadjuvant chemotherapy is useful for predicting intra-operative risk factors and the completeness of resection. Pediatr Blood Cancer 62 (9): 1543-9, 2015.
  34. Katzenstein HM, Kent PM, London WB, et al.: Treatment and outcome of 83 children with intraspinal neuroblastoma: the Pediatric Oncology Group experience. J Clin Oncol 19 (4): 1047-55, 2001.
  35. De Bernardi B, Pianca C, Pistamiglio P, et al.: Neuroblastoma with symptomatic spinal cord compression at diagnosis: treatment and results with 76 cases. J Clin Oncol 19 (1): 183-90, 2001.
  36. Simon T, Niemann CA, Hero B, et al.: Short- and long-term outcome of patients with symptoms of spinal cord compression by neuroblastoma. Dev Med Child Neurol 54 (4): 347-52, 2012.
  37. McGirt MJ, Chaichana KL, Atiba A, et al.: Incidence of spinal deformity after resection of intramedullary spinal cord tumors in children who underwent laminectomy compared with laminoplasty. J Neurosurg Pediatr 1 (1): 57-62, 2008.
  38. De Bernardi B, Quaglietta L, Haupt R, et al.: Neuroblastoma with symptomatic epidural compression in the infant: the AIEOP experience. Pediatr Blood Cancer 61 (8): 1369-75, 2014.
  39. Papathanasiou ND, Gaze MN, Sullivan K, et al.: 18F-FDG PET/CT and 123I-metaiodobenzylguanidine imaging in high-risk neuroblastoma: diagnostic comparison and survival analysis. J Nucl Med 52 (4): 519-25, 2011.
  40. Kushner BH, Kramer K, Modak S, et al.: Sensitivity of surveillance studies for detecting asymptomatic and unsuspected relapse of high-risk neuroblastoma. J Clin Oncol 27 (7): 1041-6, 2009.
  41. Owens C, Li BK, Thomas KE, et al.: Surveillance imaging and radiation exposure in the detection of relapsed neuroblastoma. Pediatr Blood Cancer 63 (10): 1786-93, 2016.
  42. Smith MA, Altekruse SF, Adamson PC, et al.: Declining childhood and adolescent cancer mortality. Cancer 120 (16): 2497-506, 2014.
  43. Corrigan JJ, Feig SA; American Academy of Pediatrics: Guidelines for pediatric cancer centers. Pediatrics 113 (6): 1833-5, 2004.

Treatment of Low-Risk Neuroblastoma

Low-risk neuroblastoma represents nearly one-half of all newly diagnosed patients. The success of previous Children's Oncology Group (COG) clinical trials has contributed to the continued reduction in therapy for select patients with neuroblastoma.

The previously used COG neuroblastoma low-risk group assignment criteria are described in Table 7.

Table 7. Children's Oncology Group (COG) Neuroblastoma Low-Risk Group Assignment Schema Used for COG Studiesa
INSS StageAgeMYCNStatusINPC ClassificationDNA Ploidyb
INPC = International Neuroblastoma Pathologic Classification; INSS = International Neuroblastoma Staging System.
a The COG-P9641 (low risk) and COG-A3961 (intermediate risk) trials established the current standard of care for neuroblastoma patients in terms of risk group assignment and treatment strategies.
b DNA Ploidy: DNA Index (DI) > 1 is favorable, = 1 is unfavorable; a hypodiploid tumor (with DI < 1) will be treated as a tumor with a DI > 1 (DI < 1 [hypodiploid] to be considered favorable ploidy).
c INSS stage 2A/2B symptomatic patients with spinal cord compression, neurologic deficits, or other symptoms are treated with immediate chemotherapy for four cycles.
d INSS stage 4S infants with favorable biology and clinical symptoms are treated with immediate chemotherapy until asymptomatic (2-4 cycles). Clinical symptoms include the following: respiratory distress with or without hepatomegaly or cord compression and neurologic deficit or inferior vena cava compression and renal ischemia; or genitourinary obstruction; or gastrointestinal obstruction and vomiting; or coagulopathy with significant clinical hemorrhage unresponsive to replacement therapy.
10-21 yAnyAnyAny
2A/2Bc<365 dAnyAnyAny
≥365 d-21 yNonamplifiedAny-
≥365 d-21 yAmplifiedFavorable-
4Sd<365 dNonamplifiedFavorable>1

Table 8 shows the International Neuroblastoma Risk Group (INRG) classification for very low-risk or low-risk neuroblastoma in use for current COG studies.

Table 8. International Neuroblastoma Risk Group (INRG) Pretreatment Classification Schema for Very Low-Risk or Low-Risk Neuroblastomaa
INRG StageHistologic CategoryGrade of Tumor DifferentiationMYCN11q AberrationPloidyPretreatment Risk Group
GN = ganglioneuroma; GNB = ganglioneuroblastoma; NA = not amplified.
a Reprinted with permission. © (2015) American Society of Clinical Oncology. All rights reserved. Pinto N et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma, J Clin Oncol 33 (27), 2015: 3008-3017.[1]
L1/L2GN maturing, GNB intermixed    A (very low)
L1Any, except GN maturing or GNB intermixed NA  B (very low)
L2 
 Age <18 moAny, except GN maturing or GNB intermixed NANo D (low)
 Age ≥18 moGNB nodular neuroblastomaDifferentiatingNANo E (low)
M 
 Age <18 mo  NA HyperdiploidF (low)
MS 
 Age <18 mo  NANo C (very low)

(Refer to the Treatment of Stage 4S Neuroblastoma section of this summary for more information about the treatment of stage 4S neuroblastoma.)

Treatment Options for Low-Risk Neuroblastoma

For patients with localized disease that appears to be resectable (either based on the absence of image-defined risk factors [L1] or on the surgeon's expertise), the tumor should be resected by an experienced surgeon. If the biology is confirmed to be favorable, residual disease is not considered a risk factor for relapse. Several studies have shown that patients with favorable biology and residual disease have excellent outcomes, with event-free survival (EFS) exceeding 90% and overall survival (OS) ranging from 99% to 100%.[2,3]

Treatment options for low-risk neuroblastoma include the following:

  1. Surgery followed by observation.
  2. Chemotherapy with or without surgery (for symptomatic disease or unresectable progressive disease after surgery).
  3. Observation without biopsy (for perinatal neuroblastoma with small adrenal tumors).
  4. Radiation therapy (only for emergency therapy).

Surgery followed by observation

Treatment for patients categorized as low risk (refer to Table 7) may be surgery alone. Results from the COG-P9641 study showed that surgery alone, even without complete resection, can cure nearly all patients with stage 1 neuroblastoma and the vast majority of patients with asymptomatic, favorable-biology, INSS stage 2A and 2B disease.[3] Similar outcomes were seen in a nonrandomized clinical trial in Japan.[4]

Chemotherapy with or without surgery

Chemotherapy with or without surgery is used to treat symptomatic disease, unresectable progressive disease after surgery, or disease with unfavorable histology or diploid disease.

The use of chemotherapy may be restricted to specific cases such as children with MYCN-amplified stage 1 and 2 neuroblastoma and children with MYCN-nonamplified stage 2B neuroblastoma who are older than 18 months or who have unfavorable histology or diploid disease. These children have a less favorable outcome than do other low-risk patients.[3,5]

Chemotherapy is also reserved for low-risk patients who are symptomatic (e.g., spinal cord compression or, in stage 4S, respiratory compromise secondary to hepatic infiltration). The chemotherapy consists of carboplatin, cyclophosphamide, doxorubicin, and etoposide. The cumulative chemotherapy dose of each agent is kept low to minimize long-term effects (COG-P9641).[3]

Evidence (chemotherapy):

  1. The COG-P9641 study was one of the first COG studies to test risk stratification based on consensus-derived factors. In this phase III nonrandomized trial, 915 patients underwent an initial operation to obtain tissue for diagnosis and biology studies and for maximal safe primary tumor resection. Chemotherapy was reserved for patients with, or at risk of, symptomatic disease, with less than 50% tumor resection at diagnosis or with unresectable progressive disease after surgery alone.[3]
    • Stage 1: Patients with stage 1 disease achieved 5-year EFS of 93% and 5-year OS of 99%.
    • Stage 2A and 2B: Asymptomatic patients with stage 2A and 2B disease (n = 306) who were observed after initial operation had a 5-year EFS of 87% and OS rate of 96%. EFS was significantly better for patients with stage 2A than for patients with 2B neuroblastoma (92% vs. 85%; P = .0321), but OS did not differ significantly (98% and 96%; P = .2867). The primary study objective (to achieve a 3-year OS of 95% for asymptomatic patients with stage 2A and 2B disease) was met. Patients with stage 2B disease had a lower EFS and OS for those with unfavorable histology (EFS, 72%; OS, 86%) or diploid tumors (EFS, 75%; OS, 84%) or for patients older than 18 months. Outcomes for patients with stage 2B, diploid tumors, and unfavorable histology were particularly poor (EFS, 54%; OS, 70%), with no survivors among the few patients with additional 1p loss of heterozygosity and all deaths occurring in children older than 18 months.
    • Asymptomatic patients at diagnosis who were observed after initial operation: Of the initial 915 patients, 800 were asymptomatic at diagnosis and observed after their initial operations. Within this group, 11% experienced recurrent or progressive disease. Of the 115 patients who received immediate chemotherapy (median, four cycles; range, one to eight), 81% of the patients had a very good partial response or better. After chemotherapy, 10% of the patients had disease recurrence or progression. For patients treated with surgery alone, the 5-year EFS rate was 89%, and the OS estimate was 97%; for patients treated with surgery and immediate chemotherapy, the 5-year EFS rate was 91%, and the overall survival estimate was 98%.
    • MYCN amplification: The impact of MYCN-amplified tumors was analyzed in patients with stage I disease. For patients with MYCN-nonamplified tumors, the 5-year EFS was 93%, and OS was 99%; for MYCN-amplified tumors, the 5-year EFS was 70% (P = .0042), and OS was 80% (P < .001).

Observation without biopsy

Observation without biopsy has been used to treat perinatal neuroblastoma with small adrenal tumors.

A COG study determined that selected small INSS stage 1 or stage 2 adrenal masses, presumed to be neuroblastoma, detected in infants younger than 6 months by screening or incidental ultrasound may safely be observed without a definitive histologic diagnosis being obtained and without surgical intervention, thus avoiding potential complications of surgery in the newborn.[6] Patients are observed frequently to detect any tumor growth or spread that would indicate a need for intervention. Additional studies, including an expansion of criteria allowing observation without surgery, are underway in the COG ANBL1232 (NCT02176967) study (refer to Table 9).

Evidence (observation without biopsy):

  1. COG-ANBL00P2 reported that expectant observation is safe in patients younger than 6 months with solid adrenal tumors smaller than 3.1 cm (or cystic tumors smaller than 5 cm) and INSS stage 1 disease, with 81% of patients demonstrating spontaneous regression while avoiding surgical intervention.[6]
    • Eighty-three of 87 eligible patients were observed without biopsy or resection; only 16 (19%) ultimately underwent surgery.
    • Three-year EFS (for a neuroblastoma event) was 97.7%, and OS was 100%.

Evidence (observation following biopsy or partial resection):

  1. Controversy exists about the need to attempt resection, whether at the time of diagnosis or later, in asymptomatic infants aged 12 months or younger with apparent stage 2B and 3 MYCN-nonamplified and favorable-biology disease. In a German clinical trial, some of these patients were observed after biopsy or partial resection without chemotherapy or radiation, and many did not progress locally and never underwent an additional resection.[7]

Treatment Options Under Clinical Evaluation

The following is an example of a national and/or institutional clinical trial that is currently being conducted. Information about ongoing clinical trials is available from the NCI website.

  • ANBL1232 (NCT02176967) (Response and Biology-Based Risk Factor-Guided Therapy in Treating Younger Patients With Non-High-Risk Neuroblastoma): This phase III trial is studying how well response and biology-based, risk factor-guided therapy works in treating younger patients with non-high-risk neuroblastoma.
Table 9. ANBL1232 Treatment Assignment for Low-Risk Neuroblastoma
INRG StageBiology (Histology and Genomicsa)AgeOtherTreatment
a Genomic features includeMYCNgene amplification, segmental chromosome aberrations (somatic copy number loss at 1p, 3p, 4p, or 11q, or somatic copy number gain at 1p, 2p, or 17q), and DNA index.
b Favorable genomic features are defined by one or more whole-chromosome gains or hyperdiploid tumor (DNA index >1) in the absence of segmental chromosome aberrations as defined above.
c Asymptomatic is defined as no life-threatening symptoms and no impending neurologic or other sequelae (e.g., epidural or intraspinal tumors with existing or impending neurologic impairment, periorbital or calvarial-based lesions with existing or impending cranial nerve impairment, or anatomic or mechanical compromise of critical organ function by tumor [abdominal compartment syndrome, urinary obstruction, etc.]).
L1 <12 months<5 cm in diameter; confirmatory study if nonadrenalObserve on study without biopsy
L2Favorable histology and genomicsb<18 monthsAsymptomaticcObserve on study
MSAny histology and genomics<3 monthsExisting or evolving hepatomegaly or symptomaticImmediate treatment, response-based chemotherapy, as per protocol
Favorable histology and genomicsb<3 monthsAsymptomaticc without existing or evolving hepatomegalyObserve per clinical scoring system
Favorable histology and genomicsb3-18 monthsAsymptomaticcObserve per clinical scoring system
SymptomaticResponse-based chemotherapy, as per protocol

Current Clinical Trials

Check the list of NCI-supported cancer clinical trials that are now accepting patients with neuroblastoma. The list of clinical trials can be further narrowed by location, drug, intervention, and other criteria.

General information about clinical trials is also available from the NCI website.

References:

  1. Pinto NR, Applebaum MA, Volchenboum SL, et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma. J Clin Oncol 33 (27): 3008-17, 2015.
  2. Matthay KK, Perez C, Seeger RC, et al.: Successful treatment of stage III neuroblastoma based on prospective biologic staging: a Children's Cancer Group study. J Clin Oncol 16 (4): 1256-64, 1998.
  3. Strother DR, London WB, Schmidt ML, et al.: Outcome after surgery alone or with restricted use of chemotherapy for patients with low-risk neuroblastoma: results of Children's Oncology Group study P9641. J Clin Oncol 30 (15): 1842-8, 2012.
  4. Iehara T, Hamazaki M, Tajiri T, et al.: Successful treatment of infants with localized neuroblastoma based on their MYCN status. Int J Clin Oncol 18 (3): 389-95, 2013.
  5. Bagatell R, Beck-Popovic M, London WB, et al.: Significance of MYCN amplification in international neuroblastoma staging system stage 1 and 2 neuroblastoma: a report from the International Neuroblastoma Risk Group database. J Clin Oncol 27 (3): 365-70, 2009.
  6. Nuchtern JG, London WB, Barnewolt CE, et al.: A prospective study of expectant observation as primary therapy for neuroblastoma in young infants: a Children's Oncology Group study. Ann Surg 256 (4): 573-80, 2012.
  7. Hero B, Simon T, Spitz R, et al.: Localized infant neuroblastomas often show spontaneous regression: results of the prospective trials NB95-S and NB97. J Clin Oncol 26 (9): 1504-10, 2008.

Treatment of Intermediate-Risk Neuroblastoma

The previously used Children's Oncology Group (COG) neuroblastoma intermediate-risk group assignment criteria are described in Table 10.

Table 10. Children's Oncology Group (COG) Neuroblastoma Intermediate-Risk Group Assignment Schema Used for the COG-A3961 Studya
INSS StageAgeMYCNStatusINPC ClassificationDNA Ploidyb
INPC = International Neuroblastoma Pathologic Classification; INSS = International Neuroblastoma Staging System.
a The COG-P9641 (low risk) and COG-A3961 (intermediate risk) trials established the current standard of care for non-high-risk neuroblastoma patients in terms of risk group assignment and treatment strategies.
b DNA Ploidy: DNA Index (DI) > 1 is favorable, DI = 1 is unfavorable; a hypodiploid tumor (with DI < 1) will be treated as a tumor with a DI > 1 (DI < 1 [hypodiploid] to be considered favorable ploidy).
c INSS stage 3 or stage 4 patients with clinical symptoms as listed above receive immediate chemotherapy.
d INSS stage 4S infants with favorable biology and clinical symptoms are treated with immediate chemotherapy until asymptomatic (2-4 cycles). Clinical symptoms include the following: respiratory distress with or without hepatomegaly or cord compression and neurologic deficit or inferior vena cava compression and renal ischemia; or genitourinary obstruction; or gastrointestinal obstruction and vomiting; or coagulopathy with significant clinical hemorrhage unresponsive to replacement therapy.
3c<365 dNonamplifiedAnyAny
≥365 d-21 yNonamplifiedFavorable-
4c<548 d[1,2,3]NonamplifiedAnyAny
4Sd<365 dNonamplifiedAny=1
<365 dNonamplifiedUnfavorableAny

Table 11 shows the International Neuroblastoma Risk Group (INRG) classification for intermediate-risk neuroblastoma in use for current COG studies.

Table 11. International Neuroblastoma Risk Group (INRG) Pretreatment Classification Schema for Intermediate-Risk Neuroblastomaa
INRG StageHistologic CategoryGrade of Tumor DifferentiationMYCN11q AberrationPloidyPretreatment Risk Group
GN = ganglioneuroma; GNB = ganglioneuroblastoma; NA = not amplified.
a Reprinted with permission. © (2015) American Society of Clinical Oncology. All rights reserved. Pinto N et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma, J Clin Oncol 33 (27), 2015: 3008-3017.[4]
L2 
 Age <18 moAny, except GN maturing or GNB intermixed NAYes G (intermediate)
 Age ≥18 moGNB nodular neuroblastomaPoorly differentiated or undifferentiatedNAYes H (intermediate)
NANo H (intermediate)
M 
 Age <12 mo  NA DiploidI (intermediate)
 Age 12 to <18 mo  NA DiploidJ (intermediate)

(Refer to the Treatment of Stage 4S Neuroblastoma section of this summary for more information about the treatment of stage 4S neuroblastoma.)

Treatment Options for Intermediate-Risk Neuroblastoma

Treatment options for intermediate-risk neuroblastoma include the following:

  1. Chemotherapy with or without surgery.
  2. Surgery and observation (in infants).
  3. Radiation therapy (only for emergency therapy).

Chemotherapy with or without surgery

Patients categorized as intermediate risk have been successfully treated with surgery and four to eight cycles of chemotherapy (carboplatin, cyclophosphamide, doxorubicin, and etoposide; the cumulative dose of each agent is kept low to minimize long-term effects from the chemotherapy regimen) (COG-A3961). As a rule, patients whose tumors had unfavorable biology received eight cycles of chemotherapy, compared with four cycles for patients whose tumors had favorable biology. The COG-A3961 phase III trial demonstrated that therapy could be significantly reduced for patients with intermediate-risk neuroblastoma while maintaining outstanding survival.[5] A nonrandomized clinical trial in Japan also reported excellent outcomes for infants with stage 3 neuroblastoma without MYCN amplification.[6]

Whether initial chemotherapy is indicated for all intermediate-risk infants with localized neuroblastoma requires further study.

Evidence (chemotherapy with or without surgery):

  1. In North America, the COG (COG-A3961) investigated a risk-based neuroblastoma treatment plan that assigned all patients to a low-, intermediate-, or high-risk group based on age, International Neuroblastoma Staging System (INSS) stage, and tumor biology (i.e., MYCN gene amplification, International Neuroblastoma Pathology Classification system, and DNA ploidy).

    This study investigated an overall reduction in treatment compared with previous treatment plans in patients with unresectable, localized, MYCN-nonamplified tumors and infants with stage 4 MYCN -nonamplified disease. The intermediate-risk group received four to eight cycles of moderate-dose neoadjuvant chemotherapy (carboplatin, cyclophosphamide, doxorubicin, and etoposide), additional surgery in some instances, and avoided radiation therapy. Of the 464 patients with intermediate-risk tumors (stages 3, 4, and 4S), 69.6% had favorable features, defined as hyperdiploidy and favorable histology, and were assigned to receive four cycles of chemotherapy.[5]

    • The administration of neoadjuvant chemotherapy facilitated at least a partial resection of 99.6% of the previously unresectable tumors. No significant difference was noted in overall survival (OS) according to the degree of resection accomplished (complete vs. incomplete, P = .37).
    • Only 2.5% of the 479 patients received local radiation therapy. The 3-year event-free survival (EFS) was 88%, and OS was 95%.
    • The 3-year EFS was 92% for patients with stage 3 disease with favorable histopathology (n = 269); 90% for patients with stage 4S disease and unfavorable biology, including diploidy or unfavorable histology (n = 31); and 81% for infants with stage 4 disease (n = 176) (P < .001 for stages 3 and 4S vs. stage 4).
    • Only infants were stratified by ploidy; those with diploid tumors received eight versus four cycles of chemotherapy. The 3-year OS estimates were 98% for stage 3 disease, 97% for stage 4S disease, and 93% for stage 4 disease (P = .002 for stages 3 and 4S vs. stage 4). Infants with diploidy had a poorer outcome (P = .03), as did all patients with diploidy studied, combined (P = .03).
    • There was no difference in OS in patients with favorable biologic features between those who received eight cycles of chemotherapy (100%) for persistent disease and those who received four cycles (96%).
    • There was no unexpected toxicity.
  2. A German prospective clinical trial enrolled 340 infants aged 1 year or younger whose tumors were stage 1, 2, or 3, histologically verified, and lacked MYCN amplification. Chemotherapy was given at diagnosis to 57 infants with organs threatened by tumor. The tumor was completely resected or nearly so in 190 infants who underwent low-risk surgery. A total of 93 infants whose tumors were not resectable without high-risk surgery, because of age or organ involvement, were observed without chemotherapy.[7]
    • Three-year OS was excellent (95%) for infants receiving chemotherapy.
    • Further surgery was avoided in 33 infants, and chemotherapy was avoided in 72 infants.
    • The 3-year OS rate for the infants who were observed without treatment was 99%. The metastases-free survival rate was 94% for infants with unresected tumors and was not different from that for infants treated with surgery or chemotherapy (median follow-up, 58 months).
    • Forty-four of 93 infants with unresected tumors experienced spontaneous regression (17 were complete regressions), and 39 infants experienced progression.
    • The investigators suggested that a wait-and-see strategy is appropriate for infants with localized neuroblastoma because regressions have been observed after the first year of life.
  3. Moderate-dose chemotherapy has been shown to be effective in the prospective Infant Neuroblastoma European Study (EURO-INF-NB-STUDY-1999-99.1); about one-half of the infants with unresectable, nonmetastatic neuroblastoma and no MYCN amplification underwent a safe surgical resection and avoided long-term adverse effects.[8][Level of evidence: 3iiA]
    • The 5-year OS rate was 99%, and the EFS rate was 90% (median follow-up, 6 years).
    • In this study, infants undergoing surgical resection had a better EFS than did those who did not have surgery.
  4. A prospective International Society of Paediatric Oncology Europe Neuroblastoma (SIOPEN) trial treated children with stage 2 or stage 3 unresectable neuroblastoma and those aged 12 to 18 months, with favorable International Neuroblastoma Pathology Classification.[9][Level of evidence: 3iiD]
    • The EFS was 98% with conventional chemotherapy.
    • These results are similar to results from the COG (COG-A3961) trial.
  5. In two European prospective trials of infants with disseminated neuroblastoma without MYCN gene amplification, infants with INSS stage 3 primary or positive skeletal scintigraphy were not started on chemotherapy unless life-threatening or organ-threatening symptoms developed. When given, chemotherapy consisted of short-dose and standard-dose chemotherapy.[10]
    • OS was 100% in the 41 patients who did not have INSS stage 4S, regardless of initial chemotherapy.
    • In infants with overt metastases to the skeleton, lung, and central nervous system, the 2-year OS was 96% (n = 45).
    • No patients died of surgery-related or chemotherapy-related complications on either protocol.

In cases of abdominal neuroblastoma thought to involve the kidney, nephrectomy is not undertaken before a trial of chemotherapy has been given.[11]

Surgery and observation (in infants)

The need for chemotherapy in all asymptomatic infants with stage 3 or 4 disease is somewhat controversial, as some European studies have shown favorable outcomes with surgery and observation as described below.[10]

Evidence (surgery and observation in infants):

  1. Infants classified as stage 4 due to a primary tumor infiltrating across the midline (INSS 3 primary with metastases limited to 4S category) or positive bone scintigraphy not associated with changes in the cortical bone documented on plain radiographs and/or computed tomography were reported to have a better outcome with less aggressive chemotherapy than were other stage 4 infants (EFS, 90% vs. 27%).[12] However, a much higher proportion of those with cortical bone lesions had tumors with MYCN amplification.[12]
  2. SIOPEN conducted a prospective trial of 125 infants (n = 41 with INSS 3 primary tumors or positive scintigraphy) with disseminated neuroblastoma without MYCN amplification to see if these patients could be observed in the absence of symptoms. However, treating physicians did not always follow the wait-and-see strategy.[10]
    • There was no significant difference in 2-year OS in patients with unresectable primary tumors and patients with resectable primary tumors (97% vs. 100%) and patients with negative or with positive skeletal scintigraphy without radiologic abnormalities (100% vs. 97%).
  3. A German prospective clinical trial enrolled 340 infants aged 1 year or younger whose tumors were stage 1, 2, or 3, verified histologically, and lacked MYCN amplification. Of the 190 infants undergoing resection, 8 infants had stage 3 disease. A total of 93 infants whose tumors were not resectable without high-risk surgery, because of age or organ involvement, were observed without chemotherapy, which included 21 stage 3 patients. Fifty-seven infants, including 41 stage 3 patients, were treated with chemotherapy to control threatening symptoms.[7]
    • Three-year OS was excellent for the entire group of infants with unresected tumors (99%), infants receiving chemotherapy (95%), and infants with resected tumors (98%) (P = .45).

Radiation therapy (only for emergency therapy)

Radiation therapy for intermediate-risk patients is emergency therapy reserved for patients with the following:

  • Symptomatic life-threatening or organ-threatening tumor that does not respond rapidly enough to chemotherapy and/or surgery; and/or
  • Progressive disease.

Treatment Options Under Clinical Evaluation

The following is an example of a national and/or institutional clinical trial that is currently being conducted. Information about ongoing clinical trials is available from the NCI website.

  • ANBL1232 (NCT02176967) (Response and Biology-Based Risk Factor-Guided Therapy in Treating Younger Patients With Non-High-Risk Neuroblastoma): This phase III trial is studying how well response and biology-based, risk factor-guided therapy works in treating younger patients with non-high-risk neuroblastoma.
Table 12. ANBL1232 Treatment Assignment for Intermediate-Risk Neuroblastoma
INRG StageBiology (Histology and Genomicsa)AgeOtherTreatment
a Genomic features includeMYCNgene amplification, segmental chromosome aberrations (somatic copy number loss at 1p, 3p, 4p, or 11q, or somatic copy number gain at 1p, 2p, or 17q), and DNA index.
b Favorable genomic features are defined by one or more whole-chromosome gains or hyperdiploid tumor (DNA index >1) in the absence of segmental chromosome aberrations as defined above.
c Asymptomatic is defined as no life-threatening symptoms and no impending neurologic or other sequelae (e.g., epidural or intraspinal tumors with existing or impending neurologic impairment, periorbital or calvarial-based lesions with existing or impending cranial nerve impairment, or anatomic or mechanical compromise of critical organ function by tumor [abdominal compartment syndrome, urinary obstruction, etc.]).
d Unfavorable genomic features are defined by the presence of any segmental chromosome aberration (somatic copy number loss at 1p, 3p, 4p, or 11q, or somatic copy number gain at 1p, 2p, or 17q) or diploid tumor (DNA index = 1). This includes copy neutral loss of heterozygosity.
e Only patients withMYCN-nonamplified tumors are eligible for the ANBL1232 study.
L2Favorable histology and genomicsb<18 monthsAsymptomaticcObserve on study
MSFavorable histology and genomicsb3-18 monthsAsymptomaticcObserve per clinical scoring system
SymptomaticResponse-based chemotherapy, as per protocol
Unfavorabled /unknown histology and genomicse<18 months Response-based chemotherapy, as per protocol

Current Clinical Trials

Check the list of NCI-supported cancer clinical trials that are now accepting patients with neuroblastoma. The list of clinical trials can be further narrowed by location, drug, intervention, and other criteria.

General information about clinical trials is also available from the NCI website.

References:

  1. Schmidt ML, Lal A, Seeger RC, et al.: Favorable prognosis for patients 12 to 18 months of age with stage 4 nonamplified MYCN neuroblastoma: a Children's Cancer Group Study. J Clin Oncol 23 (27): 6474-80, 2005.
  2. London WB, Castleberry RP, Matthay KK, et al.: Evidence for an age cutoff greater than 365 days for neuroblastoma risk group stratification in the Children's Oncology Group. J Clin Oncol 23 (27): 6459-65, 2005.
  3. George RE, London WB, Cohn SL, et al.: Hyperdiploidy plus nonamplified MYCN confers a favorable prognosis in children 12 to 18 months old with disseminated neuroblastoma: a Pediatric Oncology Group study. J Clin Oncol 23 (27): 6466-73, 2005.
  4. Pinto NR, Applebaum MA, Volchenboum SL, et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma. J Clin Oncol 33 (27): 3008-17, 2015.
  5. Baker DL, Schmidt ML, Cohn SL, et al.: Outcome after reduced chemotherapy for intermediate-risk neuroblastoma. N Engl J Med 363 (14): 1313-23, 2010.
  6. Iehara T, Hamazaki M, Tajiri T, et al.: Successful treatment of infants with localized neuroblastoma based on their MYCN status. Int J Clin Oncol 18 (3): 389-95, 2013.
  7. Hero B, Simon T, Spitz R, et al.: Localized infant neuroblastomas often show spontaneous regression: results of the prospective trials NB95-S and NB97. J Clin Oncol 26 (9): 1504-10, 2008.
  8. Rubie H, De Bernardi B, Gerrard M, et al.: Excellent outcome with reduced treatment in infants with nonmetastatic and unresectable neuroblastoma without MYCN amplification: results of the prospective INES 99.1. J Clin Oncol 29 (4): 449-55, 2011.
  9. Kohler JA, Rubie H, Castel V, et al.: Treatment of children over the age of one year with unresectable localised neuroblastoma without MYCN amplification: results of the SIOPEN study. Eur J Cancer 49 (17): 3671-9, 2013.
  10. De Bernardi B, Gerrard M, Boni L, et al.: Excellent outcome with reduced treatment for infants with disseminated neuroblastoma without MYCN gene amplification. J Clin Oncol 27 (7): 1034-40, 2009.
  11. Shamberger RC, Smith EI, Joshi VV, et al.: The risk of nephrectomy during local control in abdominal neuroblastoma. J Pediatr Surg 33 (2): 161-4, 1998.
  12. Minard V, Hartmann O, Peyroulet MC, et al.: Adverse outcome of infants with metastatic neuroblastoma, MYCN amplification and/or bone lesions: results of the French society of pediatric oncology. Br J Cancer 83 (8): 973-9, 2000.

Treatment of High-Risk Neuroblastoma

The previously used Children's Oncology Group (COG) neuroblastoma high-risk group assignment criteria are described in Table 13.

Table 13. Children's Oncology Group (COG) Neuroblastoma High-Risk Group Assignment Schema
INSS StageAgeMYCNStatusINPC ClassificationDNA Ploidya
INPC = International Neuroblastoma Pathologic Classification; INSS = International Neuroblastoma Staging System.
a DNA Ploidy: DNA Index (DI) > 1 is favorable, DI = 1 is unfavorable; a hypodiploid tumor (with DI < 1) will be treated as a tumor with a DI > 1 (DI < 1 [hypodiploid] to be considered favorable ploidy).
b INSS stage 2A/2B symptomatic patients with spinal cord compression, neurologic deficits, or other symptoms are treated with immediate chemotherapy for four cycles.
c INSS stage 3 or stage 4 patients with clinical symptoms as listed above receive immediate chemotherapy.
2A/2Bb≥365 d-21 yAmplifiedUnfavorable-
3c<365 dAmplifiedAnyAny
≥365 d-21 yNonamplifiedUnfavorable-
≥365 d-21 yAmplifiedAny-
4c<365 dAmplifiedAnyAny
≥548 d-21 yAnyAny-
4S<365 dAmplifiedAnyAny

Table 14 shows the International Neuroblastoma Risk Group (INRG) classification for high-risk neuroblastoma in use for current COG studies.

Table 14. International Neuroblastoma Risk Group (INRG) Pretreatment Classification Schema for High-Risk Neuroblastomaa
INRG StageHistologic CategoryGrade of Tumor DifferentiationMYCN11q AberrationPloidyPretreatment Risk Group
GN = ganglioneuroma; GNB = ganglioneuroblastoma.
a Reprinted with permission. © (2015) American Society of Clinical Oncology. All rights reserved. Pinto N et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma, J Clin Oncol 33 (27), 2015: 3008-3017.[1]
L1Any, except GN maturing or GNB intermixed Amplified  K (high)
L2 
 Age ≥18 moGNB nodular neuroblastomaPoorly differentiated or undifferentiatedAmplified  N (high)
M 
 Age <18 mo  Amplified  O (high)
 Age ≥18 mo     P (high)
MS
 Age <18 mo  NAYes Q (high)
Amplified  R (high)

Approximately 8% to 10% of infants with stage 4S disease will have MYCN-amplified tumors and are usually treated on high-risk protocols. The overall event-free survival (EFS) and overall survival (OS) for infants with stage 4 and 4S disease and MYCN-amplification were only 30% at 2 to 5 years after treatment in a European study.[2]

For children with high-risk neuroblastoma, long-term survival with current treatments is about 54%.[3] Children with aggressively treated, high-risk neuroblastoma may develop late recurrences, some more than 5 years after completion of therapy.[4,5]

A study from the International Neuroblastoma Risk Group database found 146 patients with distant metastases limited to lymph nodes, termed stage 4N, who tended to have favorable-biology disease and a good outcome (5-year OS, 85%), which suggests that for this special subgroup of high-risk, stage 4 patients, less-intensive therapy might be considered.[6]

Treatment Options for High-Risk Neuroblastoma

Outcomes for patients with high-risk neuroblastoma remain poor despite recent improvements in survival in randomized trials.

Treatment options for high-risk neuroblastoma typically include the following:

  1. A regimen of chemotherapy, surgery, myeloablative therapy and stem cell transplant (SCT), radiation therapy, and dinutuximab with interleukin-2 (IL-2)/granulocyte-macrophage colony-stimulating factor (GM-CSF) and isotretinoin.

Chemotherapy, surgery, myeloablative therapy and SCT, radiation therapy, and dinutuximab, with IL-2/GM-CSF and isotretinoin

Treatment for patients with high-risk disease is generally divided into the following three phases:

  • Induction (includes chemotherapy and surgical resection).
  • Consolidation (myeloablative therapy and SCT and radiation therapy to the site of the primary tumor and residual metastatic sites).
  • Postconsolidation (immunotherapy and retinoid).

Induction phase

The backbone of the most commonly used induction therapy includes dose-intensive cycles of cisplatin and etoposide alternating with vincristine, cyclophosphamide, and doxorubicin.[7] Topotecan was added to this regimen on the basis of the antineuroblastoma activity seen in relapsed patients.[8] Response to therapy after four cycles of chemotherapy or at the end of induction chemotherapy correlates with EFS at the completion of high-risk therapy.[9,10]

After a response to chemotherapy, resection of the primary tumor is usually attempted. Whether a gross-total resection is beneficial either before or after induction chemotherapy is controversial.[11]

  • The COG A3973 (NCT00004188) study had central surgical review of 220 patients undergoing attempted gross-total resection after induction chemotherapy. Degree of resection was determined to be 90% or greater versus less than 90% by the surgeon's estimate, but only 63% concordance with central review of imaging was found. Nevertheless, the surgeon's assessment of 90% or greater versus less than 90% predicted EFS of 46% versus 38% (P = .01) and cumulative incidence of local relapse of 8.5% versus 20%. OS was not significantly different (49% vs. 57%, P = .3). The author's conclusion supports continued efforts to achieve greater than 90% resection in order to decrease local recurrence.[12][Level of evidence: 3iiA]
  • A single-center retrospective study of 87 children with high-risk neuroblastoma demonstrated no significant benefit of gross-total resection compared with near-total (>90%) resection.[13][Level of evidence: 3iiD] However, the results suggest that greater than 90% resection is associated with improved OS compared with less than 90% resection.

Consolidation phase

The consolidation phase of high-risk regimens involves myeloablative chemotherapy and SCT, which attempts to eradicate minimal residual disease using lethal doses of chemotherapy and autologous stem cells collected during induction chemotherapy to repopulate the bone marrow. Several large randomized controlled studies have shown an improvement in 3-year EFS for SCT (31% to 47%) versus conventional chemotherapy (22% to 31%).[14,15,16] Previously, total-body irradiation had been used in SCT conditioning regimens. Most current protocols use either carboplatin/etoposide/melphalan or busulfan/melphalan as conditioning for SCT.[17][Level of evidence: 3iA] Two sequential cycles of myeloablative chemotherapy and stem cell rescue given in a tandem fashion has been shown to be feasible for patients with high-risk neuroblastoma.[18]

A randomized clinical study (COG-ANBL0532) testing the efficacy of two cycles versus one cycle of myeloablative chemotherapy with stem cell rescue has been completed.[19] Children older than 18 months with stage 4 neuroblastoma who had received six cycles of induction chemotherapy were then randomly assigned to receive a single autologous SCT with carboplatin/etoposide/melphalan or tandem transplants with cyclophosphamide/thiotepa followed by reduced-dose carboplatin/etoposide/melphalan. After tumor bed radiation, patients were randomly assigned on a second trial to receive isotretinoin alone or isotretinoin with dinutuximab. The 3-year EFS was 61% for tandem transplants and 48% for single autologous SCT (P = .008). The 3-year OS was 74% for tandem autologous SCTs and 69% for single autologous SCT (P = .19). For patients who were randomly assigned to not receive dinutuximab, the 3-year EFS was 74% for tandem SCTs and 56% for single autologous SCT (P = .003); for patients who received dinutuximab, the 3-year OS was 84% for tandem SCTs and 74% for single SCT (P = .03). (Refer to the Autologous Hematopoietic Cell Transplantation section in the PDQ summary on Childhood Hematopoietic Cell Transplantation for more information about transplantation.)

Radiation to the primary tumor site (whether or not a complete excision was obtained) and persistently mIBG-positive bony metastatic sites is often performed after myeloablative therapy. The optimal dose of radiation therapy has not been determined. Radiation of metastatic disease sites is determined on an individual basis or according to protocol guidelines for patients enrolled in studies.

Metastatic bone relapse in neuroblastoma usually occurs at anatomic sites of previous disease. Metastatic sites identified at diagnosis that did not receive radiation during frontline therapy appeared to have a higher risk of involvement at first relapse relative to previously irradiated metastatic sites.[20] These observations support the current paradigm of irradiating metastases that persist after induction chemotherapy in high-risk patients.

Preliminary outcomes for proton radiation therapy of high-risk neuroblastoma primary tumors have been published.[21]

Postconsolidation phase

Postconsolidation therapy is designed to treat potential minimal residual disease following SCT.[22] For high-risk patients in remission after SCT, dinutuximab combined with GM-CSF and IL-2 are given in concert with isotretinoin and have been shown to improve EFS.[23,24]

Evidence (all treatments):

  1. A randomized study was performed comparing high-dose therapy with purged autologous bone marrow transplant (ABMT) versus three cycles of intensive consolidation chemotherapy. In addition, after the completion of either chemotherapy or ABMT, patients on this study were randomly assigned to stop therapy or to receive 6 months of isotretinoin.[14]; [22][Level of evidence: 1iiA] The EFS and OS results described below reflect outcome from the time of each randomization.
    • The 5-year EFS was significantly better in the ABMT arm (30%), than in the consolidation chemotherapy arm (19%; P = .04). There was no significant difference in 5-year OS between the two arms (39% vs. 30%; P = .08).[22]
    • Patients who received isotretinoin had a higher 5-year EFS than did patients who received no maintenance therapy (42% vs. 31%), although the difference was not significant (P = .12). OS was higher for patients randomly assigned to receive isotretinoin (50%) than it was for those who stopped therapy (39%), but this difference was not significant (P = .10).[22]
  2. An updated Cochrane review evaluated three randomized clinical trials comparing ABMT with standard chemotherapy.[14,15,16,22,25]
    • EFS was significantly better for ABMT, but there was no statistically significant difference in OS.
  3. A retrospective, single-institution, nonrandomized trial compared patients who received GM-CSF and 3F8 anti-GD2 antibody therapy after either autologous SCT or conventional chemotherapy.[26] The patients were a mixture of those referred for initial treatment or further therapy, and included refractory and relapsed patients, some of whom had received autologous SCT at referring institutions. In the autologous SCT group, there was a significantly longer time from first chemotherapy or from autologous SCT to initiation of GM-CSF and 3F8 anti-GD2 antibody treatment. The autologous SCT group also had significantly more ultra-high-risk patients.
    • A trend for better EFS with GM-CSF and 3F8 anti-GD2 antibody therapy and autologous SCT was observed (65% vs. 51%, P = .128), but there was no statistically significant difference in OS between those treated with chemotherapy alone and those treated with autologous SCT.
  4. In a separate prospective, randomized study, there was no advantage to purging harvested stem cells of neuroblastoma cells before transplantation.[27]
  5. A review of 147 allogeneic transplant cases submitted to the Center for International Blood and Marrow Transplant Research found no advantage for allogeneic transplant over autologous transplant, even if the allogeneic transplant recipient had received a previous autologous transplant.[28]
  6. In a COG phase III trial after SCT, patients were randomly assigned to receive dinutuximab administered with GM-CSF and IL-2 in conjunction with isotretinoin, versus isotretinoin alone.[23] Dinutuximab has been approved by the U.S. Food and Drug Administration.
    • Immunotherapy together with isotretinoin (EFS, 66%) was superior to standard isotretinoin maintenance therapy (EFS, 46%). As a result, immunotherapy post-SCT is considered the standard of care in COG trials for high-risk disease.

Surgery and radiation therapy (local control)

The potential benefit of aggressive surgical approaches in high-risk patients with metastatic disease to achieve complete tumor resection, either at the time of diagnosis or after chemotherapy, has not been unequivocally demonstrated.

  • Several studies have reported that complete resection of the primary tumor at diagnosis improved survival; however, the outcome in these patients may be more dependent on the biology of the tumor, which itself may determine resectability, than on the extent of surgical resection.[29,30,31]
  • Radiation therapy to consolidate local control after surgical resection is often given.[32]; [33][Level of evidence: 3iiA]
  • In stage 4 patients older than 18 months, controversy exists about whether there is any advantage to gross-total resection of the primary tumor after chemotherapy.[12,30,31,34]

Current Clinical Trials

Check the list of NCI-supported cancer clinical trials that are now accepting patients with neuroblastoma. The list of clinical trials can be further narrowed by location, drug, intervention, and other criteria.

General information about clinical trials is also available from the NCI website.

References:

  1. Pinto NR, Applebaum MA, Volchenboum SL, et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma. J Clin Oncol 33 (27): 3008-17, 2015.
  2. Canete A, Gerrard M, Rubie H, et al.: Poor survival for infants with MYCN-amplified metastatic neuroblastoma despite intensified treatment: the International Society of Paediatric Oncology European Neuroblastoma Experience. J Clin Oncol 27 (7): 1014-9, 2009.
  3. Maris JM: Recent advances in neuroblastoma. N Engl J Med 362 (23): 2202-11, 2010.
  4. Cotterill SJ, Pearson AD, Pritchard J, et al.: Late relapse and prognosis for neuroblastoma patients surviving 5 years or more: a report from the European Neuroblastoma Study Group "Survey". Med Pediatr Oncol 36 (1): 235-8, 2001.
  5. Mertens AC, Yasui Y, Neglia JP, et al.: Late mortality experience in five-year survivors of childhood and adolescent cancer: the Childhood Cancer Survivor Study. J Clin Oncol 19 (13): 3163-72, 2001.
  6. Morgenstern DA, London WB, Stephens D, et al.: Metastatic neuroblastoma confined to distant lymph nodes (stage 4N) predicts outcome in patients with stage 4 disease: A study from the International Neuroblastoma Risk Group Database. J Clin Oncol 32 (12): 1228-35, 2014.
  7. Kushner BH, LaQuaglia MP, Bonilla MA, et al.: Highly effective induction therapy for stage 4 neuroblastoma in children over 1 year of age. J Clin Oncol 12 (12): 2607-13, 1994.
  8. Park JR, Scott JR, Stewart CF, et al.: Pilot induction regimen incorporating pharmacokinetically guided topotecan for treatment of newly diagnosed high-risk neuroblastoma: a Children's Oncology Group study. J Clin Oncol 29 (33): 4351-7, 2011.
  9. Decarolis B, Schneider C, Hero B, et al.: Iodine-123 metaiodobenzylguanidine scintigraphy scoring allows prediction of outcome in patients with stage 4 neuroblastoma: results of the Cologne interscore comparison study. J Clin Oncol 31 (7): 944-51, 2013.
  10. Yanik GA, Parisi MT, Shulkin BL, et al.: Semiquantitative mIBG scoring as a prognostic indicator in patients with stage 4 neuroblastoma: a report from the Children's oncology group. J Nucl Med 54 (4): 541-8, 2013.
  11. De Ioris MA, Crocoli A, Contoli B, et al.: Local control in metastatic neuroblastoma in children over 1 year of age. BMC Cancer 15: 79, 2015.
  12. von Allmen D, Davidoff AM, London WB, et al.: Impact of Extent of Resection on Local Control and Survival in Patients From the COG A3973 Study With High-Risk Neuroblastoma. J Clin Oncol 35 (2): 208-216, 2017.
  13. Englum BR, Rialon KL, Speicher PJ, et al.: Value of surgical resection in children with high-risk neuroblastoma. Pediatr Blood Cancer 62 (9): 1529-35, 2015.
  14. Matthay KK, Villablanca JG, Seeger RC, et al.: Treatment of high-risk neuroblastoma with intensive chemotherapy, radiotherapy, autologous bone marrow transplantation, and 13-cis-retinoic acid. Children's Cancer Group. N Engl J Med 341 (16): 1165-73, 1999.
  15. Berthold F, Boos J, Burdach S, et al.: Myeloablative megatherapy with autologous stem-cell rescue versus oral maintenance chemotherapy as consolidation treatment in patients with high-risk neuroblastoma: a randomised controlled trial. Lancet Oncol 6 (9): 649-58, 2005.
  16. Pritchard J, Cotterill SJ, Germond SM, et al.: High dose melphalan in the treatment of advanced neuroblastoma: results of a randomised trial (ENSG-1) by the European Neuroblastoma Study Group. Pediatr Blood Cancer 44 (4): 348-57, 2005.
  17. Elborai Y, Hafez H, Moussa EA, et al.: Comparison of toxicity following different conditioning regimens (busulfan/melphalan and carboplatin/etoposide/melphalan) for advanced stage neuroblastoma: Experience of two transplant centers. Pediatr Transplant 20 (2): 284-9, 2016.
  18. Seif AE, Naranjo A, Baker DL, et al.: A pilot study of tandem high-dose chemotherapy with stem cell rescue as consolidation for high-risk neuroblastoma: Children's Oncology Group study ANBL00P1. Bone Marrow Transplant 48 (7): 947-52, 2013.
  19. Park JR, Kreissman SG, London WB, et al.: A phase III randomized clinical trial (RCT) of tandem myeloablative autologous stem cell transplant (ASCT) using peripheral blood stem cell (PBSC) as consolidation therapy for high-risk neuroblastoma (HR-NB): a Children's Oncology Group (COG) study. [Abstract] J Clin Oncol 34 (Suppl 15): A-LBA3, 2016. Also available online. Last accessed November 07, 2016.
  20. Polishchuk AL, Li R, Hill-Kayser C, et al.: Likelihood of bone recurrence in prior sites of metastasis in patients with high-risk neuroblastoma. Int J Radiat Oncol Biol Phys 89 (4): 839-45, 2014.
  21. Hattangadi JA, Rombi B, Yock TI, et al.: Proton radiotherapy for high-risk pediatric neuroblastoma: early outcomes and dose comparison. Int J Radiat Oncol Biol Phys 83 (3): 1015-22, 2012.
  22. Matthay KK, Reynolds CP, Seeger RC, et al.: Long-term results for children with high-risk neuroblastoma treated on a randomized trial of myeloablative therapy followed by 13-cis-retinoic acid: a children's oncology group study. J Clin Oncol 27 (7): 1007-13, 2009.
  23. Yu AL, Gilman AL, Ozkaynak MF, et al.: Anti-GD2 antibody with GM-CSF, interleukin-2, and isotretinoin for neuroblastoma. N Engl J Med 363 (14): 1324-34, 2010.
  24. Cheung NK, Cheung IY, Kushner BH, et al.: Murine anti-GD2 monoclonal antibody 3F8 combined with granulocyte-macrophage colony-stimulating factor and 13-cis-retinoic acid in high-risk patients with stage 4 neuroblastoma in first remission. J Clin Oncol 30 (26): 3264-70, 2012.
  25. Yalçin B, Kremer LC, Caron HN, et al.: High-dose chemotherapy and autologous haematopoietic stem cell rescue for children with high-risk neuroblastoma. Cochrane Database Syst Rev 8: CD006301, 2013.
  26. Kushner BH, Ostrovnaya I, Cheung IY, et al.: Lack of survival advantage with autologous stem-cell transplantation in high-risk neuroblastoma consolidated by anti-GD2 immunotherapy and isotretinoin. Oncotarget 7 (4): 4155-66, 2016.
  27. Kreissman SG, Seeger RC, Matthay KK, et al.: Purged versus non-purged peripheral blood stem-cell transplantation for high-risk neuroblastoma (COG A3973): a randomised phase 3 trial. Lancet Oncol 14 (10): 999-1008, 2013.
  28. Hale GA, Arora M, Ahn KW, et al.: Allogeneic hematopoietic cell transplantation for neuroblastoma: the CIBMTR experience. Bone Marrow Transplant 48 (8): 1056-64, 2013.
  29. DeCou JM, Bowman LC, Rao BN, et al.: Infants with metastatic neuroblastoma have improved survival with resection of the primary tumor. J Pediatr Surg 30 (7): 937-40; discussion 940-1, 1995.
  30. Castel V, Tovar JA, Costa E, et al.: The role of surgery in stage IV neuroblastoma. J Pediatr Surg 37 (11): 1574-8, 2002.
  31. Simon T, Häberle B, Hero B, et al.: Role of surgery in the treatment of patients with stage 4 neuroblastoma age 18 months or older at diagnosis. J Clin Oncol 31 (6): 752-8, 2013.
  32. Haas-Kogan DA, Swift PS, Selch M, et al.: Impact of radiotherapy for high-risk neuroblastoma: a Children's Cancer Group study. Int J Radiat Oncol Biol Phys 56 (1): 28-39, 2003.
  33. Gatcombe HG, Marcus RB Jr, Katzenstein HM, et al.: Excellent local control from radiation therapy for high-risk neuroblastoma. Int J Radiat Oncol Biol Phys 74 (5): 1549-54, 2009.
  34. Adkins ES, Sawin R, Gerbing RB, et al.: Efficacy of complete resection for high-risk neuroblastoma: a Children's Cancer Group study. J Pediatr Surg 39 (6): 931-6, 2004.

Treatment of Stage 4S Neuroblastoma

Many patients with stage 4S neuroblastoma do not require therapy. However, tumors with unfavorable biology or patients who are symptomatic due to evolving hepatomegaly and organ compromise are at increased risk of death and are treated with low-dose to moderate-dose chemotherapy. Eight percent to 10% of these patients will have MYCN amplification and are treated with high-risk protocols.[1]

The previously used Children's Oncology Group (COG) neuroblastoma 4S group assignment criteria are described in Table 15.

Table 15. Children's Oncology Group (COG) Neuroblastoma Stage 4S Group Assignment Schema Used for COG-P9641, COG-A3961, and COG-A3973 Studiesa
INSS StageAgeMYCNStatusINPC ClassificationDNA PloidybRisk Group
INPC = International Neuroblastoma Pathologic Classification; INSS = International Neuroblastoma Staging System.
a The COG-P9641, COG-A3961, and COG-A3973 trials established the current standard of care for neuroblastoma patients in terms of risk group assignment and treatment strategies.
b DNA Ploidy: DNA Index (DI) > 1 is favorable, = 1 is unfavorable; a hypodiploid tumor (with DI < 1) will be treated as a tumor with a DI > 1 (DI < 1 [hypodiploid] to be considered favorable ploidy).
c INSS stage 4S infants with favorable biology and clinical symptoms are treated with immediate chemotherapy until asymptomatic or according to protocol guidelines. Clinical symptoms include the following: respiratory distress with or without hepatomegaly or cord compression and neurologic deficit or inferior vena cava compression and renal ischemia; or genitourinary obstruction; or gastrointestinal obstruction and vomiting; or coagulopathy with significant clinical hemorrhage unresponsive to replacement therapy.
4Sc<365 dNonamplifiedFavorable>1Low
<365 dNonamplifiedAny=1Intermediate
<365 dNonamplifiedUnfavorableAnyIntermediate
<365 dAmplifiedAnyAnyHigh

Table 16 shows the International Neuroblastoma Risk Group (INRG) classification for stage 4S neuroblastoma in use for current COG studies.

Table 16. International Neuroblastoma Risk Group (INRG) Pretreatment Classification Schema for Stage 4S Neuroblastomaa
INRG StageHistologic CategoryGrade of Tumor DifferentiationMYCN11q AberrationPloidyPretreatment Risk Group
NA = not amplified.
a Reprinted with permission. © (2015) American Society of Clinical Oncology. All rights reserved. Pinto N et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma, J Clin Oncol 33 (27), 2015: 3008-3017.[2]
MS 
 Age <18 mo  NANo C (very low)
Yes Q (high)
Amplified  R (high)

Treatment Options for Stage 4S Neuroblastoma

There is no standard approach to the treatment of stage 4S neuroblastoma.

Treatment options for stage 4S neuroblastoma include the following:

  1. Observation with supportive care (for asymptomatic patients with favorable tumor biology).
  2. Chemotherapy (for symptomatic patients, very young infants, or those with unfavorable biology).

Resection of primary tumor is not associated with improved outcome.[3,4,5] Rarely, infants with massive hepatic 4S neuroblastoma develop cirrhosis from the chemotherapy and/or radiation therapy that is used to control the disease and may benefit from orthotopic liver transplantation.[6]

Observation with supportive care

Observation with supportive care is used to treat asymptomatic patients with favorable tumor biology.

The treatment of children with stage 4S disease is dependent on clinical presentation.[3,4] Most patients do not require therapy unless bulk disease is causing organ compromise and risk of death.

Chemotherapy

Chemotherapy is used to treat symptomatic patients, very young infants, or those with unfavorable biology.

Infants diagnosed with International Neuroblastoma Staging System (INSS) stage 4S neuroblastoma, particularly those with hepatomegaly or those younger than 2 months, have the potential for rapid clinical deterioration and may benefit from early initiation of therapy. It has been difficult to identify infants with stage 4S disease who will benefit from chemotherapy. Several clinical trials have evaluated the presence of symptoms in patients with 4S disease, including the following:

  • Of 45 patients with stage 4S neuroblastoma diagnosed in the first month of life, 16 patients developed dyspnea caused by massive liver enlargement; one-half of them did not survive.[7]
  • A review of 35 patients with INSS stage 4S disease described 13 patients younger than 4 weeks, all of whom had liver involvement. Of the seven who died, all presented with hepatomegaly at birth, and all deaths resulted from hepatomegaly or related complications. Of the infants who were aged 1 month to 12 months (n = 22), 21 had hepatomegaly, and there were three deaths (14%). Deaths were due to infection, disseminated intravascular coagulation, and radiation nephritis. One death was related to hepatomegaly.

    A scoring system to measure signs and symptoms of deterioration or compromise was developed to better assess this group.[8] This scoring system has been evaluated retrospectively, was predictive of the clinical course, and has been applied prospectively. It was also helpful in directing the management of patients with INSS stage 4S disease.[8,9]

Various chemotherapy regimens (cyclophosphamide alone, carboplatin/etoposide, cyclophosphamide/doxorubicin/vincristine) have been used to treat symptomatic patients. The approach is to administer the chemotherapy only as long as symptoms persist to avoid toxicity, which contributes to poorer survival. Additionally, lower doses of chemotherapy are often recommended for very young or low-weight infants, along with granulocyte colony-stimulating factors after each cycle of chemotherapy.

Evidence (chemotherapy for symptomatic patients, very young infants, or those with unfavorable biology):

  1. Eighty stage 4S patients were enrolled on COG-P9641.[10]
    • Overall, the 5-year event-free survival (EFS) was 77%, and the overall survival (OS) was 91%.
    • The 5-year EFS was 63% and OS was 84% for the 41 patients with asymptomatic stage 4S neuroblastoma treated with surgery alone; the EFS was 95% and OS was 97% for the 39 patients treated with surgery and chemotherapy (EFS P = .0016; OS P = .1302). Previously, chemotherapy toxicity was thought to be responsible for the poorer survival of patients with stage 4S disease; however, the use of chemotherapy on COG-P9641 was restricted to specific clinical situations with a recommended number of cycles.
  2. Also, on COG-P9641, asymptomatic infants with biologically favorable (MYCN-nonamplified) INSS stage 4S disease did not receive chemotherapy until the development of progressive disease or clinical symptoms.[10]
    • Infants who became symptomatic had disease-related organ failure and infectious complications resulting in an inferior OS compared with those who received immediate chemotherapy (four to eight cycles of therapy). The 3-year OS for infants who did not receive chemotherapy was 84% versus 97% for infants who received chemotherapy (P = .1321).
  3. On COG-ANBL0531, the 2-year OS rate for INSS stage 4S patients was 81%, which is lower than that reported on COG-P9641 and is thought to reflect the expanded eligibility allowing enrollment of patients who were too ill to undergo diagnostic biopsy. These patients would have been excluded from previous COG trials.[11]
  4. A prospective study was performed in 125 infants with stage 4S MYCN-nonamplified tumors or INSS stage 3 primary tumors and/or positive bone scintigraphy not associated with changes in the cortical bone documented on plain radiographs and/or CT. A pretreatment symptom score was used to determine initial treatment; observation was recommended for infants with low symptom scores (n = 86) and chemotherapy for infants with high symptom scores (n = 37). The chemotherapy recommended for patients with high symptom scores included two to four 3-day courses of carboplatin and etoposide; if symptoms persisted or progressive disease developed, up to four 5-day courses of cyclophosphamide, doxorubicin, and vincristine were administered. One-half of the patients underwent complete or partial resection of the primary tumor.[9]
    • There was no difference in the 2-year EFS and OS between asymptomatic and symptomatic patients (EFS, 87% vs. 88%; OS, 98% vs. 97%), although many of the investigators preferred to give chemotherapy in the presence of a low symptom score.
    • For infants with low symptom scores, there was no difference between the outcome in the initially untreated infants (n = 56; OS, 93%) and treated infants (n = 30; OS, 86%).
    • The OS was 90% for infants presenting with high symptom scores.
    • There was no significant difference in 2-year OS in patients with unresectable primary tumors and patients with resectable primary tumors (97% vs. 100%) and patients with negative or with positive skeletal scintigraphy without radiologic abnormalities (100% vs. 97%).

Treatment Options Under Clinical Evaluation

The following is an example of a national and/or institutional clinical trial that is currently being conducted. Information about ongoing clinical trials is available from the NCI website.

  • ANBL1232 (NCT02176967) (Response and Biology-Based Risk Factor-Guided Therapy in Treating Younger Patients With Non-High-Risk Neuroblastoma):
    • For all newly diagnosed INRG MS patients younger than 18 months, the following occurs:
      • Patients younger than 3 months with existing or evolving hepatomegaly or who are symptomatic are entered onto the trial, and chemotherapy begins immediately. Full staging must be completed within 1 month; a tumor biopsy is not performed until the patient is stable.
      • Patients aged 3 to 12 months who are symptomatic are entered onto the trial, and chemotherapy begins immediately. Tumor biopsy is performed after the patient is stable.
      • Patients aged 12 to 18 months who are symptomatic have a tumor biopsy before starting chemotherapy.
      • Patients aged 3 to 18 months who are asymptomatic and patients younger than 3 months who are asymptomatic and have no evolving hepatomegaly have a tumor biopsy followed by close observation initially, to continue for 3 years.

        Patients with INRG MS tumors that have unfavorable histology or unfavorable genomic features with or without symptoms are treated according to a response-based algorithm to determine length of treatment. For INRG MS patients under observation without chemotherapy, an objective scoring system is used to monitor them for clinical changes and initiate therapy. For patients with complete resolution of symptoms and at least a 50% reduction in primary tumor volume (partial response), chemotherapy is discontinued, and observation continues for 3 years after completion of therapy. If the disease progresses, the patient leaves this study.

References:

  1. Canete A, Gerrard M, Rubie H, et al.: Poor survival for infants with MYCN-amplified metastatic neuroblastoma despite intensified treatment: the International Society of Paediatric Oncology European Neuroblastoma Experience. J Clin Oncol 27 (7): 1014-9, 2009.
  2. Pinto NR, Applebaum MA, Volchenboum SL, et al.: Advances in Risk Classification and Treatment Strategies for Neuroblastoma. J Clin Oncol 33 (27): 3008-17, 2015.
  3. Guglielmi M, De Bernardi B, Rizzo A, et al.: Resection of primary tumor at diagnosis in stage IV-S neuroblastoma: does it affect the clinical course? J Clin Oncol 14 (5): 1537-44, 1996.
  4. Katzenstein HM, Bowman LC, Brodeur GM, et al.: Prognostic significance of age, MYCN oncogene amplification, tumor cell ploidy, and histology in 110 infants with stage D(S) neuroblastoma: the pediatric oncology group experience--a pediatric oncology group study. J Clin Oncol 16 (6): 2007-17, 1998.
  5. Nickerson HJ, Matthay KK, Seeger RC, et al.: Favorable biology and outcome of stage IV-S neuroblastoma with supportive care or minimal therapy: a Children's Cancer Group study. J Clin Oncol 18 (3): 477-86, 2000.
  6. Steele M, Jones NL, Ng V, et al.: Successful liver transplantation in an infant with stage 4S(M) neuroblastoma. Pediatr Blood Cancer 60 (3): 515-7, 2013.
  7. Gigliotti AR, Di Cataldo A, Sorrentino S, et al.: Neuroblastoma in the newborn. A study of the Italian Neuroblastoma Registry. Eur J Cancer 45 (18): 3220-7, 2009.
  8. Hsu LL, Evans AE, D'Angio GJ: Hepatomegaly in neuroblastoma stage 4s: criteria for treatment of the vulnerable neonate. Med Pediatr Oncol 27 (6): 521-8, 1996.
  9. De Bernardi B, Gerrard M, Boni L, et al.: Excellent outcome with reduced treatment for infants with disseminated neuroblastoma without MYCN gene amplification. J Clin Oncol 27 (7): 1034-40, 2009.
  10. Strother DR, London WB, Schmidt ML, et al.: Outcome after surgery alone or with restricted use of chemotherapy for patients with low-risk neuroblastoma: results of Children's Oncology Group study P9641. J Clin Oncol 30 (15): 1842-8, 2012.
  11. Park JR, Bagatell R, London WB, et al.: Children's Oncology Group's 2013 blueprint for research: neuroblastoma. Pediatr Blood Cancer 60 (6): 985-93, 2013.

Recurrent Neuroblastoma

Tumor growth resulting from maturation should be differentiated from tumor progression by performing a biopsy and reviewing histology. Patients may have persistent maturing disease with metaiodobenzylguanidine (mIBG) uptake that does not affect outcome, particularly patients with low-risk and intermediate-risk disease.[1] An analysis of 23 paired mIBG and positron emission tomography (PET) scans in 14 patients with refractory or recurrent high-risk neuroblastoma treated with iodine-131 mIBG (131 I-mIBG) found that mIBG was more sensitive than fludeoxyglucose F-18 (FDG)-PET for detecting metastatic bone lesions, although there was a trend for FDG-PET to be more sensitive for soft tissue lesions.[2]

Subclonal ALK mutations or other MAPK pathway lesions may be present at diagnosis, with subsequent clonal expansion at relapse. Consequently, serial sampling of progressive tumors may lead to the identification of potential actionable mutations.[3] Modern comprehensive molecular analysis comparing primary and relapsed neuroblastoma from the same patients revealed extensive clonal enrichment and several newly discovered mutations, with many tumors showing new or clonal-enriched mutations in the RAS-MAPK pathway. This was true for patients with both high-risk and low-risk tumors at diagnosis.[4,5]

If neuroblastoma recurs in a child originally diagnosed with high-risk disease, the prognosis is usually poor despite additional intensive therapy.[6,7,8,9] However, it is often possible to gain many additional months of life for these patients with alternative chemotherapy regimens.[10,11] Clinical trials are appropriate for these patients and may be offered. Information about ongoing clinical trials is available from the NCI website.

Prognostic Factors for Recurrent Neuroblastoma

The International Neuroblastoma Risk Group Project performed a survival-tree analysis of clinical and biological characteristics (defined at diagnosis) associated with survival after relapse in 2,266 patients with neuroblastoma entered on large clinical trials in well-established clinical trials groups around the world.[6]

  • Overall survival (OS) in the entire relapse population was 20%.
  • Among patients with all stages of disease at diagnosis, MYCN amplification predicted a poorer prognosis, measured as 5-year OS.
  • Among patients diagnosed with International Neuroblastoma Staging System (INSS) stage 4 without amplification, age older than18 months and high lactate dehydrogenase (LDH) level predicted poor prognosis.
  • Among patients with MYCN amplification, those diagnosed with stages 1 and 2 have a better prognosis than do those diagnosed with stages 3 and 4.
  • Among patients with MYCN-nonamplified who are not stage 4, patients with hyperdiploidy had a better prognosis than did patients with diploidy in those younger than 18 months, while among those older than 18 months, patients with differentiating tumors did much better than did patients with undifferentiated and poorly differentiated tumors.

Significant prognostic factors determined at diagnosis for postrelapse survival include the following:[6]

  • Age.
  • INSS stage.
  • MYCN status.
  • Time from diagnosis to first relapse.
  • LDH level, ploidy, and histologic grade of tumor differentiation (to a lesser extent).

Although the OS after recurrence in children presenting with high-risk neuroblastoma is generally extremely poor, patients with high-risk neuroblastoma at first relapse after complete remission or minimal residual disease in whom relapse was a single site of soft tissue mass (a few children also had bone marrow or bone disease at relapse) had a 5-year OS of 35% in one single-institution study. All patients underwent surgical resection of the soft tissue disease. MYCN amplification and multifocal soft tissue disease were associated with a worse postprogression survival.[12]

The Children's Oncology Group (COG) experience with recurrence in patients with low-risk and intermediate-risk neuroblastoma is that most patients can be salvaged. The COG reported a 3-year event free survival (EFS) of 88% and an OS of 96% in intermediate-risk patients and a 5-year EFS of 89% and OS of 97% in low-risk patients.[13,14] Moreover, in most patients originally diagnosed with low-risk or intermediate-risk disease, local recurrence or recurrence in the 4S pattern may be treated successfully with surgery and/or with moderate dose chemotherapy, without myeloablative therapy and stem cell transplantation.

Recurrent Neuroblastoma in Patients Initially Classified as Low Risk

Locoregional recurrence

Treatment options for locoregional recurrent neuroblastoma initially classified as low risk include the following:

  1. Surgery followed by observation or chemotherapy.
  2. Chemotherapy that may be followed by surgery.

Local or regional recurrent cancer is resected if possible.

Those with favorable biology and regional recurrence more than 3 months after completion of planned treatment are observed if resection of the recurrence is total or near total (≥90% resection). Those with favorable biology and a less-than-near-total resection are treated with chemotherapy.

Infants younger than 1 year at the time of locoregional recurrence whose tumors have any unfavorable biologic properties are observed if resection is total or near total. If the resection is less than near total, these same infants are treated with chemotherapy. Chemotherapy may consist of moderate doses of carboplatin, cyclophosphamide, doxorubicin, and etoposide, or cyclophosphamide and topotecan. The cumulative dose of each agent is kept low to minimize long-term effects from the chemotherapy regimen as used in previous COG trials (COG-P9641 and COG-A3961).

Older children with local recurrence with either unfavorable International Neuroblastoma Pathology Classification at diagnosis or MYCN gene amplification have a poor prognosis and may be treated with surgery, aggressive combination chemotherapy, or they may be offered entry into a clinical trial.

Evidence (surgery followed by observation or chemotherapy):

  1. A COG study of treatment of low-risk patients with stage 1, 2A, 2B, and 4S neuroblastoma enrolled 915 patients, 800 of whom were asymptomatic and were treated with surgery alone followed by observation. The others received chemotherapy with or without surgery.[14]
    • About 10% of patients developed progressive or recurrent tumor. Most recurrences were treated on study with surgery alone or moderate chemotherapy with or without surgery, and most patients were salvaged as demonstrated by the EFS (89%) and OS (97%) rates at 5 years.

Metastatic recurrence

Treatment options for metastatic recurrent neuroblastoma initially classified as low risk include the following:

  1. Observation.
  2. Chemotherapy.
  3. Surgery followed by chemotherapy.
  4. High-risk therapy.

Metastatic recurrent or progressive neuroblastoma in an infant initially categorized as low risk and younger than 1 year at recurrence may be treated according to tumor biology as defined in the previous COG trials (COG-P9641 and COG-A3961):

  1. If the biology is completely favorable, metastasis is in a 4S pattern, and the recurrence or progression is within 3 months of diagnosis, the patient is observed systematically.
  2. If the metastatic progression or recurrence occurs more than 3 months after diagnosis or not in a 4S pattern, then the primary tumor is resected, if possible, and chemotherapy is given.

    Chemotherapy may consist of moderate doses of carboplatin, cyclophosphamide, doxorubicin, and etoposide. The cumulative dose of each agent is kept low to minimize long-term effects from the chemotherapy regimen, as used in previous COG trials (COG-P9641 and COG-A3961).

Any child initially categorized as low risk who is older than 1 year at the time of metastatic recurrent or progressive disease and whose recurrence is not in the stage 4S pattern usually has a poor prognosis and is treated as follows:

  1. High-risk therapy.

Patients with metastatic recurrent neuroblastoma are treated like patients with newly diagnosed high-risk neuroblastoma. (Refer to the Treatment Options for High-Risk Neuroblastoma section of this summary for more information.)

Recurrent Neuroblastoma in Patients Initially Classified as Intermediate Risk

The treatment options for locoregional and metastatic recurrence in patients with intermediate-risk neuroblastoma are derived from the results of the COG-A3961 trial. Among 479 patients with intermediate-risk neuroblastoma treated on the COG-A3961 clinical trial, 42 patients developed disease progression. The rate was 10% of those with favorable biology and 17% of those with unfavorable biology. Thirty patients had locoregional recurrence, 11 had metastatic recurrence, and 1 had both types of recurrent disease. Six of the 42 patients died of disease, while 36 patients were salvaged. Thus, most patients with intermediate-risk neuroblastoma and disease progression may be salvaged.[13]

Locoregional recurrence

Treatment options for locoregional recurrent neuroblastoma initially classified as intermediate risk include the following:

  1. Surgery (complete resection).
  2. Surgery (incomplete resection) followed by chemotherapy.

The current standard of care is based on the experience from the COG Intermediate-Risk treatment plan (COG-A3961). Locoregional recurrence of neuroblastoma with favorable biology that occurs more than 3 months after completion of chemotherapy may be treated surgically. If resection is less than near total, then additional chemotherapy may be given. Chemotherapy may consist of moderate doses of carboplatin, cyclophosphamide, doxorubicin, and etoposide. The cumulative dose of each agent is kept low to minimize long-term effects from the chemotherapy regimen, as used in a previous COG trial (COG-A3961).

Metastatic recurrence

Treatment options for metastatic recurrent neuroblastoma initially classified as intermediate risk include the following:

  1. High-risk therapy.

Patients with metastatic recurrent neuroblastoma are treated like patients with newly diagnosed high-risk neuroblastoma. (Refer to the Treatment Options for High-Risk Neuroblastoma section of this summary for more information.)

Recurrent Neuroblastoma in Patients Initially Classified as High Risk

Any recurrence in patients initially classified as high risk signifies a very poor prognosis.[6] Clinical trials may be considered. Palliative care should also be considered as part of the patient's treatment plan.

Although the OS after recurrence in children presenting with high-risk neuroblastoma is generally extremely poor, patients with high-risk neuroblastoma at first relapse after complete remission or minimal residual disease in whom relapse was a single site of soft tissue mass (a few children also had bone marrow or bone disease at relapse) had a 5-year OS of 35% in one single-institution study.[12]

Treatment options for recurrent or refractory neuroblastoma in patients initially classified as high risk include the following:

  1. Chemotherapy.
    • Topotecan in combination with cyclophosphamide or etoposide.[15]
    • Temozolomide with irinotecan.
  2. 131 I-mIBG alone, in combination with other therapy, or followed by stem cell rescue.
  3. Second autologous stem cell transplantation (SCT) after additional chemotherapy. (Refer to the Autologous Hematopoietic Cell Transplantation section in the PDQ summary on Childhood Hematopoietic Cell Transplantation for more information about transplantation.)
  4. Crizotinib, or other ALK inhibitors, for patients with ALK mutations.[16]

It is not known whether one therapeutic approach is superior to another.

Evidence (chemotherapy with or without autologous SCT):

  1. Topotecan in combination with cyclophosphamide or etoposide has been used in patients with recurrent disease who did not receive topotecan initially.[17,18]; [15][Level of evidence: 1A]
  2. The combination of irinotecan and temozolomide had a 15% response rate in one study.[19][Level of evidence: 2A]
  3. High-dose carboplatin, irinotecan, and/or temozolomide has been used in relapsed patients resistant or refractory to regimens containing topotecan.[18]
  4. A retrospective study reported on 74 patients who received 92 cycles of ifosfamide, carboplatin, and etoposide; it included 37 patients who received peripheral blood stem cell rescue after responding to this drug combination.[20]
    • Disease regressions (major and minor responses) were achieved by 14 of 17 patients (82%) with a new relapse, 13 of 26 patients (50%) with refractory neuroblastoma, and 12 of 34 patients (35%) who were treated for progressive disease during chemotherapy (P = .005).
    • Grade 3 toxicities were rare.
  5. Children with high-risk neuroblastoma refractory after induction chemotherapy (N = 26), defined as poor response of metastatic disease, underwent tandem autologous SCT with thiotepa followed by busulfan/melphalan. Data were prospectively recorded in the Gustave Roussy Pediatric ASCT database.[21]
    • The 3-year EFS was 37%.

Evidence (131 I-mIBG):

  1. For children with recurrent or refractory neuroblastoma, 131 I-mIBG is an effective palliative agent and may be considered alone or in combination with chemotherapy (with stem cell rescue) in a clinical research trial.[22,23,24,25,26,27]; [28,29][Level of evidence: 3iiiA]
  2. A North American retrospective study of more than 200 patients treated with 131 I-mIBG therapy compared children who had recurrence or progression of disease with children who had stable or persistent disease since diagnosis.[30]
    • The rate of immediate progression after mIBG therapy was lower and OS at 2 years was better (65% vs. 39%) in those with stable, persistent disease.
  3. Tandem consolidation using 131 I-mIBG, vincristine, and irinotecan with autologous SCT followed by busulfan/melphalan with autologous SCT was retrospectively reported in eight patients and resulted in three complete responses, two partial responses, and one minor response.[29]
  4. Single autologous SCT with escalating dose 131 I-mIBG and carboplatin/etoposide/melphalan was studied in additional patients.[31]
    • After induction chemotherapy, 27 refractory patients and 15 progressing patients were treated, resulting in four responses. Eight patients with partial response to induction were treated, resulting in three responses.
    • The 12% incidence of sinusoidal obstructive syndrome was dose limiting.

Evidence (second autologous SCT following additional chemotherapy):

  1. Second autologous SCT after additional chemotherapy may be considered, particularly in the setting of a clinical trial. (Refer to the Autologous Hematopoietic Cell Transplantation section in the PDQ summary on Childhood Hematopoietic Cell Transplantation for more information about transplantation.)
  2. Data from three consecutive German high-risk neuroblastoma trials described 253 children relapsing after intensive chemotherapy with autologous SCT and who had a 5-year OS rate lower than 10%. Only 23 of the 253 patients eventually proceeded to a second autologous SCT after additional chemotherapy.[32][Level of evidence: 3iiiA]
    • Among these patients, the 3-year OS rate was 43%, but the 5-year OS rate was lower than 20%.
    • This shows that intensive second-line therapy is feasible, although even with intensive therapy and second autologous SCT, only a small minority of relapsed high-risk neuroblastoma patients may benefit.

Allogeneic transplantation has a historically low success rate in recurrent or progressive neuroblastoma. In a retrospective registry study, allogeneic SCT after a previous autologous SCT appeared to offer minimal benefit. Disease recurrence remains the most common cause of treatment failure.[33]

Clinical trials of novel therapeutic approaches, such as a vaccine designed to induce host antiganglioside antibodies that can replicate the antineoplastic activities of intravenously administered monoclonal antibodies, are currently under investigation. Patients also receive a beta-glucan treatment, which has a broad range of immunostimulatory effects and synergizes with anti-GD2/GD3 monoclonal antibodies. In a phase I study of 15 children with high-risk neuroblastoma, the therapy was tolerated without any dose-limiting toxicity.[34] Long-term progression-free survival (PFS) has been reported in patients who achieve a second or later complete or very good partial remission followed by consolidation with anti-GD2 immunotherapy and isotretinoin with or without maintenance therapy. This includes patients who had previously received anti-GD2 immunotherapy and isotretinoin.[35]

Recurrent Neuroblastoma in the Central Nervous System

Central nervous system (CNS) involvement, although rare at initial presentation, may occur in 5% to 10% of patients with recurrent neuroblastoma. Because upfront treatment for newly diagnosed patients does not adequately treat the CNS, the CNS has emerged as a sanctuary site leading to relapse.[36,37] CNS relapses are almost always fatal, with a median time to death of 6 months.

Treatment options for recurrent neuroblastoma in the CNS include the following:

  1. Surgery and radiation therapy.
  2. Novel therapeutic approaches.

Current treatment approaches generally include eradicating bulky and microscopic residual disease in the CNS and minimal residual systemic disease that may herald further relapses. Neurosurgical interventions serve to decrease edema, control hemorrhage, and remove bulky tumor before starting therapy.

Compartmental radioimmunotherapy using intrathecal radioiodinated monoclonal antibodies has been tested in patients with recurrent metastatic CNS neuroblastoma after surgery, craniospinal radiation therapy, and chemotherapy.[11]

Treatment Options Under Clinical Evaluation for Recurrent or Refractory Neuroblastoma

The following are examples of national and/or institutional clinical trials that are currently being conducted. Information about ongoing clinical trials is available from the NCI website.

  • COG-ANBL1221 (NCI-2012-03125; NCT01767194) (A Phase II Randomized Trial of Irinotecan/Temozolomide with Temsirolimus or Chimeric 14.18 Antibody [ch14.18] in Children with Refractory, Relapsed, or Progressive Neuroblastoma): This Pick the Winner phase II study is designed to compare the response rates and PFS for patients with refractory, relapsed, or progressive neuroblastoma receiving temsirolimus or ch14.18 in combination with irinotecan and temozolomide. Patients older than 365 days who have progressed from INSS stage 1, 2, or 4S and have received no chemotherapy or only one cycle of chemotherapy are eligible for this trial. (Refer to the featured clinical trial, Monoclonal Antibody Therapy for Relapsed or Treatment-Resistant Neuroblastoma, for more information.)
  • NANT N2011-04 (NCI-2012-02011; NCT01711554) (Lenalidomide and Monoclonal Antibody With or Without Isotretinoin in Treating Younger Patients With Refractory or Recurrent Neuroblastoma): This study aims to determine the maximum tolerated dose and/or recommended phase II dose of lenalidomide in combination with fixed doses of ch14.18 given intravenously for 4 days (days 8-11) and isotretinoin given twice each day orally for 14 days (days 15-28) and repeated every 28 days for children with refractory or recurrent neuroblastoma. (Refer to the featured clinical trial, Monoclonal Antibody Therapy for Relapsed or Treatment-Resistant Neuroblastoma, for more information.)
  • NANT N2011-01 (NCT02035137) (131 I-mIBG Alone vs. 131 I-mIBG With Vincristine and Irinotecan vs. 131 I-MIBG With Vorinostat): The New Approaches to Neuroblastoma Therapy (NANT) consortium developed a randomized, selection-design phase II trial evaluating 131 I-mIBG alone versus 131 I-mIBG plus vorinostat or vincristine and irinotecan in patients with relapsed, refractory, or persistent neuroblastoma.
  • NCT00911560 (Bivalent Vaccine With Escalating Doses of the Immunological Adjuvant OPT-821, in Combination With Oral Beta-Glucan for High-Risk Neuroblastoma): The purpose of this study is to test the safety of a vaccine against neuroblastoma and its effect on cancer.
  • ADVL1212 (NCT01606878) (Crizotinib and Combination Chemotherapy in Treating Younger Patients With Relapsed or Refractory Solid Tumors or Anaplastic Large Cell Lymphoma): This is a phase I study of the ALK inhibitor crizotinib in combination with conventional chemotherapy for relapsed or refractory solid tumors or anaplastic large cell lymphoma.
  • ADVL1522 (NCT02452554) (Lorvotuzumab Mertansine in Treating Younger Patients with Relapsed or Refractory Wilms Tumor, Rhabdomyosarcoma, Neuroblastoma, Pleuropulmonary Blastoma, Malignant Peripheral Nerve Sheath Tumor, or Synovial Sarcoma): This is a phase II study of IMGN901 (lorvotuzumab mertansine) in children with relapsed or refractory Wilms tumor, rhabdomyosarcoma, neuroblastoma, pleuropulmonary blastoma, malignant peripheral nerve sheath tumor and synovial sarcoma. This trial is studying the effects of the IMGN901 antibody-drug conjugate that links a potent anti-mitotic to antibodies that target CD56.

Current Clinical Trials

Check the list of NCI-supported cancer clinical trials that are now accepting patients with recurrent neuroblastoma. The list of clinical trials can be further narrowed by location, drug, intervention, and other criteria.

General information about clinical trials is also available from the NCI website.

References:

  1. Marachelian A, Shimada H, Sano H, et al.: The significance of serial histopathology in a residual mass for outcome of intermediate risk stage 3 neuroblastoma. Pediatr Blood Cancer 58 (5): 675-81, 2012.
  2. Taggart DR, Han MM, Quach A, et al.: Comparison of iodine-123 metaiodobenzylguanidine (MIBG) scan and [18F]fluorodeoxyglucose positron emission tomography to evaluate response after iodine-131 MIBG therapy for relapsed neuroblastoma. J Clin Oncol 27 (32): 5343-9, 2009.
  3. Schleiermacher G, Javanmardi N, Bernard V, et al.: Emergence of new ALK mutations at relapse of neuroblastoma. J Clin Oncol 32 (25): 2727-34, 2014.
  4. Eleveld TF, Oldridge DA, Bernard V, et al.: Relapsed neuroblastomas show frequent RAS-MAPK pathway mutations. Nat Genet 47 (8): 864-71, 2015.
  5. Schramm A, Köster J, Assenov Y, et al.: Mutational dynamics between primary and relapse neuroblastomas. Nat Genet 47 (8): 872-7, 2015.
  6. London WB, Castel V, Monclair T, et al.: Clinical and biologic features predictive of survival after relapse of neuroblastoma: a report from the International Neuroblastoma Risk Group project. J Clin Oncol 29 (24): 3286-92, 2011.
  7. Pole JG, Casper J, Elfenbein G, et al.: High-dose chemoradiotherapy supported by marrow infusions for advanced neuroblastoma: a Pediatric Oncology Group study. J Clin Oncol 9 (1): 152-8, 1991.
  8. Castel V, Cañete A, Melero C, et al.: Results of the cooperative protocol (N-III-95) for metastatic relapses and refractory neuroblastoma. Med Pediatr Oncol 35 (6): 724-6, 2000.
  9. Lau L, Tai D, Weitzman S, et al.: Factors influencing survival in children with recurrent neuroblastoma. J Pediatr Hematol Oncol 26 (4): 227-32, 2004.
  10. Saylors RL 3rd, Stine KC, Sullivan J, et al.: Cyclophosphamide plus topotecan in children with recurrent or refractory solid tumors: a Pediatric Oncology Group phase II study. J Clin Oncol 19 (15): 3463-9, 2001.
  11. Kramer K, Kushner BH, Modak S, et al.: Compartmental intrathecal radioimmunotherapy: results for treatment for metastatic CNS neuroblastoma. J Neurooncol 97 (3): 409-18, 2010.
  12. Murphy JM, Lim II, Farber BA, et al.: Salvage rates after progression of high-risk neuroblastoma with a soft tissue mass. J Pediatr Surg 51 (2): 285-8, 2016.
  13. Baker DL, Schmidt ML, Cohn SL, et al.: Outcome after reduced chemotherapy for intermediate-risk neuroblastoma. N Engl J Med 363 (14): 1313-23, 2010.
  14. Strother DR, London WB, Schmidt ML, et al.: Outcome after surgery alone or with restricted use of chemotherapy for patients with low-risk neuroblastoma: results of Children's Oncology Group study P9641. J Clin Oncol 30 (15): 1842-8, 2012.
  15. London WB, Frantz CN, Campbell LA, et al.: Phase II randomized comparison of topotecan plus cyclophosphamide versus topotecan alone in children with recurrent or refractory neuroblastoma: a Children's Oncology Group study. J Clin Oncol 28 (24): 3808-15, 2010.
  16. Mossé YP, Lim MS, Voss SD, et al.: Safety and activity of crizotinib for paediatric patients with refractory solid tumours or anaplastic large-cell lymphoma: a Children's Oncology Group phase 1 consortium study. Lancet Oncol 14 (6): 472-80, 2013.
  17. Simon T, Längler A, Harnischmacher U, et al.: Topotecan, cyclophosphamide, and etoposide (TCE) in the treatment of high-risk neuroblastoma. Results of a phase-II trial. J Cancer Res Clin Oncol 133 (9): 653-61, 2007.
  18. Kushner BH, Kramer K, Modak S, et al.: Differential impact of high-dose cyclophosphamide, topotecan, and vincristine in clinical subsets of patients with chemoresistant neuroblastoma. Cancer 116 (12): 3054-60, 2010.
  19. Bagatell R, London WB, Wagner LM, et al.: Phase II study of irinotecan and temozolomide in children with relapsed or refractory neuroblastoma: a Children's Oncology Group study. J Clin Oncol 29 (2): 208-13, 2011.
  20. Kushner BH, Modak S, Kramer K, et al.: Ifosfamide, carboplatin, and etoposide for neuroblastoma: a high-dose salvage regimen and review of the literature. Cancer 119 (3): 665-71, 2013.
  21. Pasqualini C, Dufour C, Goma G, et al.: Tandem high-dose chemotherapy with thiotepa and busulfan-melphalan and autologous stem cell transplantation in very high-risk neuroblastoma patients. Bone Marrow Transplant 51 (2): 227-31, 2016.
  22. DuBois SG, Groshen S, Park JR, et al.: Phase I Study of Vorinostat as a Radiation Sensitizer with 131I-Metaiodobenzylguanidine (131I-MIBG) for Patients with Relapsed or Refractory Neuroblastoma. Clin Cancer Res 21 (12): 2715-21, 2015.
  23. Polishchuk AL, Dubois SG, Haas-Kogan D, et al.: Response, survival, and toxicity after iodine-131-metaiodobenzylguanidine therapy for neuroblastoma in preadolescents, adolescents, and adults. Cancer 117 (18): 4286-93, 2011.
  24. Matthay KK, Yanik G, Messina J, et al.: Phase II study on the effect of disease sites, age, and prior therapy on response to iodine-131-metaiodobenzylguanidine therapy in refractory neuroblastoma. J Clin Oncol 25 (9): 1054-60, 2007.
  25. Matthay KK, Tan JC, Villablanca JG, et al.: Phase I dose escalation of iodine-131-metaiodobenzylguanidine with myeloablative chemotherapy and autologous stem-cell transplantation in refractory neuroblastoma: a new approaches to Neuroblastoma Therapy Consortium Study. J Clin Oncol 24 (3): 500-6, 2006.
  26. Matthay KK, Quach A, Huberty J, et al.: Iodine-131--metaiodobenzylguanidine double infusion with autologous stem-cell rescue for neuroblastoma: a new approaches to neuroblastoma therapy phase I study. J Clin Oncol 27 (7): 1020-5, 2009.
  27. DuBois SG, Chesler L, Groshen S, et al.: Phase I study of vincristine, irinotecan, and ¹³¹I-metaiodobenzylguanidine for patients with relapsed or refractory neuroblastoma: a new approaches to neuroblastoma therapy trial. Clin Cancer Res 18 (9): 2679-86, 2012.
  28. Johnson K, McGlynn B, Saggio J, et al.: Safety and efficacy of tandem 131I-metaiodobenzylguanidine infusions in relapsed/refractory neuroblastoma. Pediatr Blood Cancer 57 (7): 1124-9, 2011.
  29. French S, DuBois SG, Horn B, et al.: 131I-MIBG followed by consolidation with busulfan, melphalan and autologous stem cell transplantation for refractory neuroblastoma. Pediatr Blood Cancer 60 (5): 879-84, 2013.
  30. Zhou MJ, Doral MY, DuBois SG, et al.: Different outcomes for relapsed versus refractory neuroblastoma after therapy with (131)I-metaiodobenzylguanidine ((131)I-MIBG). Eur J Cancer 51 (16): 2465-72, 2015.
  31. Yanik GA, Villablanca JG, Maris JM, et al.: 131I-metaiodobenzylguanidine with intensive chemotherapy and autologous stem cell transplantation for high-risk neuroblastoma. A new approaches to neuroblastoma therapy (NANT) phase II study. Biol Blood Marrow Transplant 21 (4): 673-81, 2015.
  32. Simon T, Berthold F, Borkhardt A, et al.: Treatment and outcomes of patients with relapsed, high-risk neuroblastoma: results of German trials. Pediatr Blood Cancer 56 (4): 578-83, 2011.
  33. Hale GA, Arora M, Ahn KW, et al.: Allogeneic hematopoietic cell transplantation for neuroblastoma: the CIBMTR experience. Bone Marrow Transplant 48 (8): 1056-64, 2013.
  34. Kushner BH, Cheung IY, Modak S, et al.: Phase I trial of a bivalent gangliosides vaccine in combination with β-glucan for high-risk neuroblastoma in second or later remission. Clin Cancer Res 20 (5): 1375-82, 2014.
  35. Kushner BH, Ostrovnaya I, Cheung IY, et al.: Prolonged progression-free survival after consolidating second or later remissions of neuroblastoma with Anti-GD2 immunotherapy and isotretinoin: a prospective Phase II study. Oncoimmunology 4 (7): e1016704, 2015.
  36. Kramer K, Kushner B, Heller G, et al.: Neuroblastoma metastatic to the central nervous system. The Memorial Sloan-kettering Cancer Center Experience and A Literature Review. Cancer 91 (8): 1510-9, 2001.
  37. Matthay KK, Brisse H, Couanet D, et al.: Central nervous system metastases in neuroblastoma: radiologic, clinical, and biologic features in 23 patients. Cancer 98 (1): 155-65, 2003.

Changes to this Summary (04 / 14 / 2017)

The PDQ cancer information summaries are reviewed regularly and updated as new information becomes available. This section describes the latest changes made to this summary as of the date above.

General Information About Neuroblastoma

Added text about subclonal ALK mutations in neuroblastoma (cited Bellini et al. as a reference 46).

Revised text to state that in a Children's Oncology Group (COG) study investigating the effect of histology, among other factors, on outcome, 87% of 915 children with stage 1 and stage 2 neuroblastoma without MYCN amplification were treated with initial surgery and observation.

Stage Information for Neuroblastoma

Added text to state that the International Neuroblastoma Risk Group collaboration has also defined techniques for detecting and quantifying neuroblastoma in bone marrow, both at diagnosis and after treatment. Quantification of this disease may result in more accurate assessment of response to treatment, but has not yet been applied to any clinical trials (cited Burchill et al. as reference 19).

Treatment Option Overview for Neuroblastoma

Added von Allmen et al. as reference 31.

Revised text to state that in an overall surveillance plan, which includes urinary vanillylmandelic acid and homovanillic acid testing, one of the most reliable imaging tests to detect disease progression or recurrence is the 123 I-metaiodobenzylguanidine scan. Also added text to state that cross-sectional imaging with computed tomography scans is controversial because of the amount of radiation received and the low proportion of relapses detected with this modality (cited Owens et al. as reference 41).

Treatment of High-Risk Neuroblastoma

Added text to state that whether a gross-total resection is beneficial either before or after induction chemotherapy is controversial (cited De Ioris et al. as reference 11).

Added text about the results of the COG A3973 study and a single-center retrospective study on how the degree of resection affects outcome of patients with high-risk neuroblastoma (cited von Allmen et al. as reference 12 and level of evidence 3iiA and Englum et al. as reference 13 and level of evidence 3iiD).

Added Elborai et al. as reference 17 and level of evidence 3iA.

Added text about the results of a retrospective, single-institution, nonrandomized trial that compared patients who received GM-CSF and 3F8 anti-GD2 antibody therapy after either autologous stem cell transplantation or conventional chemotherapy (cited Kushner et al. as reference 26).

Recurrent Neuroblastoma

Added text about the outcomes of patients with recurrent high-risk neuroblastoma in one single-institution study (cited Murphy et al. as reference 12).

Added text about the outcomes of patients with recurrent high-risk neuroblastoma in one single-institution study.

This summary is written and maintained by the PDQ Pediatric Treatment Editorial Board, which is editorially independent of NCI. The summary reflects an independent review of the literature and does not represent a policy statement of NCI or NIH. More information about summary policies and the role of the PDQ Editorial Boards in maintaining the PDQ summaries can be found on the About This PDQ Summary and PDQ® - NCI's Comprehensive Cancer Database pages.

About This PDQ Summary

Purpose of This Summary

This PDQ cancer information summary for health professionals provides comprehensive, peer-reviewed, evidence-based information about the treatment of neuroblastoma. It is intended as a resource to inform and assist clinicians who care for cancer patients. It does not provide formal guidelines or recommendations for making health care decisions.

Reviewers and Updates

This summary is reviewed regularly and updated as necessary by the PDQ Pediatric Treatment Editorial Board, which is editorially independent of the National Cancer Institute (NCI). The summary reflects an independent review of the literature and does not represent a policy statement of NCI or the National Institutes of Health (NIH).

Board members review recently published articles each month to determine whether an article should:

  • be discussed at a meeting,
  • be cited with text, or
  • replace or update an existing article that is already cited.

Changes to the summaries are made through a consensus process in which Board members evaluate the strength of the evidence in the published articles and determine how the article should be included in the summary.

The lead reviewers for Neuroblastoma Treatment are:

  • Christopher N. Frantz, MD (Alfred I. duPont Hospital for Children)
  • Andrea A. Hayes-Jordan, MD, FACS, FAAP (M.D. Anderson Cancer Center)
  • Karen J. Marcus, MD (Dana-Farber Cancer Institute/Boston Children's Hospital)
  • Nita Louise Seibel, MD (National Cancer Institute)
  • Stephen J. Shochat, MD (St. Jude Children's Research Hospital)

Any comments or questions about the summary content should be submitted to Cancer.gov through the NCI website's Email Us. Do not contact the individual Board Members with questions or comments about the summaries. Board members will not respond to individual inquiries.

Levels of Evidence

Some of the reference citations in this summary are accompanied by a level-of-evidence designation. These designations are intended to help readers assess the strength of the evidence supporting the use of specific interventions or approaches. The PDQ Pediatric Treatment Editorial Board uses a formal evidence ranking system in developing its level-of-evidence designations.

Permission to Use This Summary

PDQ is a registered trademark. Although the content of PDQ documents can be used freely as text, it cannot be identified as an NCI PDQ cancer information summary unless it is presented in its entirety and is regularly updated. However, an author would be permitted to write a sentence such as "NCI's PDQ cancer information summary about breast cancer prevention states the risks succinctly: [include excerpt from the summary]."

The preferred citation for this PDQ summary is:

PDQ® Pediatric Treatment Editorial Board. PDQ Neuroblastoma Treatment. Bethesda, MD: National Cancer Institute. Updated <MM/DD/YYYY>. Available at: https://www.cancer.gov/types/neuroblastoma/hp/neuroblastoma-treatment-pdq. Accessed <MM/DD/YYYY>. [PMID: 26389190]

Images in this summary are used with permission of the author(s), artist, and/or publisher for use within the PDQ summaries only. Permission to use images outside the context of PDQ information must be obtained from the owner(s) and cannot be granted by the National Cancer Institute. Information about using the illustrations in this summary, along with many other cancer-related images, is available in Visuals Online, a collection of over 2,000 scientific images.

Disclaimer

Based on the strength of the available evidence, treatment options may be described as either "standard" or "under clinical evaluation." These classifications should not be used as a basis for insurance reimbursement determinations. More information on insurance coverage is available on Cancer.gov on the Managing Cancer Care page.

Contact Us

More information about contacting us or receiving help with the Cancer.gov website can be found on our Contact Us for Help page. Questions can also be submitted to Cancer.gov through the website's Email Us.

Last Revised: 2017-04-14