Pediatric Solid Tumors


Summary of Key Points

  • Osteosarcoma

  • The incidence of osteosarcoma is 5 new cases per 1 million population per year among children and adolescents (up to 19 years).

  • Osteosarcoma is the most common bone tumor in children and adolescents.

  • Other lytic bone lesions, including eosinophilic granuloma and giant cell tumor, must be excluded.

  • At histopathological examination, tumors must be differentiated from fibrosarcoma and chondrosarcoma.

  • Staging evaluation includes a complete history and physical examination, complete blood cell count (CBC), serum chemistry analysis (including determination of alkaline phosphatase level), and imaging studies of the primary tumor and chest.

  • The primary tumor is managed surgically with a limb-sparing operation or amputation.

  • Chemotherapy with high-dose methotrexate, cisplatin, and doxorubicin is the standard of care. For patients receiving preoperative chemotherapy, good histologic response is associated with a greater than 80% chance of cure.

  • Primary pulmonary metastatic disease is managed surgically, although adjuvant chemotherapy may reduce the extent of resection.

  • Pulmonary metastatic disease is managed surgically.

  • Postthoracotomy adjuvant chemotherapy is of unproven benefit.

  • Local recurrence is managed surgically.

  • Further adjuvant chemotherapy with non–cross-resistant agents is of uncertain benefit.

  • Ewing Sarcoma Family Tumors

  • The incidence of Ewing sarcoma family tumors is 2.9 new cases per 1 million population per year among children and adolescents (0–19 years).

  • Ewing sarcoma is the second most common bone tumor among children and adolescents.

  • Osteomyelitis must be ruled out, especially when the patient has a fever.

  • Lesions to be excluded include benign tumors of bone that manifest as lytic lesions (e.g., eosinophilic granuloma, giant cell tumor), malignant tumors (e.g., osteosarcoma, primary lymphoma of bone), and metastatic lesions from a nonosseous tumor (e.g., neuroblastoma).

  • Staging evaluation includes a complete history and physical examination, CBC, serum chemistry analysis (including determination of lactate dehydrogenase level), bone marrow biopsy, and imaging studies of the primary tumor, bones, and chest.

  • The primary tumor should always be treated with multimodality therapy consisting of chemotherapy, radiation therapy, surgery, or a combination of these treatments.

  • Specific local treatment depends on the primary site. Surgical extirpation may be considered with tumors in expendable bones (proximal part of the fibula, rib, clavicle, iliac wing).

  • Unresectable tumors generally necessitate a combined approach of chemotherapy, radiation therapy, and surgery.

  • Localized disease is curable with combined therapy in more than 70% of cases. Metastatic disease is curable in 30% to 40% of cases.

  • Effective second-line therapy has not been established.

  • Neuroblastoma

  • Among children younger than 19 years of age, 8.2 new cases occur per 1 million population per year.

  • Neuroblastoma is the most common extracranial solid tumor in children.

  • Disseminated bone disease can resemble systemic infection, inflammatory disease, osteomyelitis, or rheumatoid arthritis.

  • Paraneoplastic syndromes associated with neuroblastoma (vasoactive intestinal peptide syndrome, opsoclonus-myoclonus-ataxia syndrome) must be differentiated from primary inflammatory bowel disease and neurologic disease.

  • Differential diagnosis must include other small, round, blue cell undifferentiated neoplasms of childhood (e.g., Ewing sarcoma, primitive neuroectodermal tumor, rhabdomyosarcoma, non-Hodgkin lymphoma, undifferentiated soft tissue sarcoma).

  • Approximately 40% of tumors have amplification of the proto-oncogene MYCN. In as many as 10% of tumors, catecholamines are not produced. In 1%, the absence of an obvious primary lesion confounds the diagnosis.

  • Staging evaluation includes a complete history and physical examination, CBC, serum chemistry analysis (including determination of lactate dehydrogenase level), quantitative urine catecholamines, bone marrow examination, radionuclide scintigraphy, and imaging studies of the primary tumor and chest.

  • Neuroblastoma is a tumor derived from the sympathetic nervous system, and neuroblastoma cells typically express the norepinephrine receptor. Thus radiolabeled m -iodobenzylguanidine (MIBG) scintigraphy is highly sensitive and specific for imaging both primary and metastatic disease.

  • Specific treatment depends on stage of disease, age of the patient, histology, and biological features of the tumor.

  • With few exceptions (completely resected primary tumor, localized tumor with/without complete excision and favorable tumor biological features, infants with stable stage 4S disease), multiple-agent chemotherapy is the backbone of multimodality treatment.

  • Survival depends on stage and biological features of the tumor (e.g., histopathological subtype, MYCN gene amplification, and other genomic alterations).

  • The overall survival (OS) rate for asymptomatic patients with favorable biology stage 1, 2a, or 2b tumors is greater than 90%.

  • The OS rate among patients with advanced-stage 3 or 4 with MYCN gene amplification disease is poor. Fewer than half of children with high-risk disease are long-term survivors.

  • No second-line therapy has proved to increase survival.

  • Wilms Tumor

  • The incidence of Wilms tumor is 7.8 cases per 1 million population per year among children younger than 15 years of age.

  • The mean ages at diagnosis are 44 months for unilateral tumors and 31 months for bilateral tumors.

  • Familial cases account for 1.5% of cases of Wilms tumor.

  • Associated with Wilms tumor is a syndrome comprising aniridia, genitourinary anomalies, and mental retardation (WAGR syndrome); Denys-Drash syndrome; and Beckwith-Wiedemann syndrome.

  • No firmly established environmental factors have been identified.

  • The main histologic subtypes are favorable and anaplastic.

  • Major implicated genes include WT1 , IGF2 , FAM123B/WTX , CTNNB1 /β-catenin, and DROSHA , DGCR8 , SIX1/2 , and MLLT1 .

  • Mutations of the p53 gene (TP53) are associated with anaplastic tumors.

  • Loss of heterozygosity at chromosomes 1p and 16q and gain of chromosome 1q are associated with adverse prognosis.

  • An asymptomatic abdominal mass is found in most patients.

  • Other characteristic features include abdominal pain, hematuria, hypertension, and congenital anomalies (genitourinary malformations, aniridia, and hemihypertrophy).

  • Neuroblastoma is the main consideration in the differential diagnosis.

  • Other tumors to be excluded are renal neoplasms (clear cell sarcoma, rhabdoid tumor, congenital mesoblastic nephroma, renal cell carcinoma).

  • Benign renal processes (nephrogenic rests, multicystic or polycystic kidneys, hydronephrosis, renal carbuncles, hemorrhage) also should be ruled out.

  • Evaluation begins with complete history and physical examination, with careful attention to blood pressure and assessment for associated congenital anomalies.

  • Other components of the staging evaluation include CBC, serum chemistry analysis, urinalysis, abdominal ultrasonography, abdominal or pelvic computed tomography (CT) or magnetic resonance imaging (MRI), and chest CT.

  • Surgery: Surgical resection of the primary tumor usually precedes chemotherapy in the approach taken by the Children's Oncology Group (COG). Presurgical chemotherapy is used in the approach taken by the International Society of Pediatric Oncology (SIOP).

  • Chemotherapy: Agents used depend on disease stage and favorable versus anaplastic histology: stages I and II, favorable histology—vincristine, actinomycin D; stages III and IV, favorable histology, and stage I, anaplastic histology—vincristine, doxorubicin, actinomycin D; and stages II through IV, anaplastic histology—vincristine, cyclophosphamide, doxorubicin, carboplatin, etoposide.

  • Radiation therapy: Radiation therapy is included in management for stages III and IV, favorable histology, and stages I through IV, anaplastic histology.

  • Recurrent disease is effectively managed with radiation therapy and chemotherapy with agents not used for initial treatment.

  • Patients who initially received aggressive treatment may respond to cyclophosphamide or ifosfamide-, carboplatin-, and etoposide-based regimens.

  • Renal failure is seen in fewer than 1% of patients with unilateral tumors and 12% in patients with bilateral tumors.

  • Congestive heart failure may occur in patients who receive doxorubicin (occurs in 4.4% of patients).

  • Pregnancy-related complications in adulthood have occurred in girls who receive flank irradiation.

  • A second malignant neoplasm also may develop (6.7% cumulative incidence of solid malignancies at age 40 years).

  • With favorable histology, outcomes include 85% 4-year relapse-free survival rate and 90% 4-year OS rate.

  • With anaplastic histology, outcomes include 50% 4-year relapse-free survival rate and 50% 4-year OS rate.

  • Renal Cell Carcinoma

  • The incidence of renal cell carcinoma is 7 cases per 10 million population per year among children younger than 20 years of age.

  • Mean age at diagnosis in pediatric series is 9 years.

  • Other renal tumors of childhood, primarily Wilms tumor, should be ruled out.

  • Staging evaluation includes a complete history and physical examination, CBC, serum chemistry analysis, and imaging studies of the abdomen, pelvis, and chest.

  • Nephrectomy of the involved kidney constitutes definitive therapy for primary tumors.

  • For metastatic disease, enrollment in clinical studies of investigational agents is recommended. Tyrosine kinase inhibitors have shown benefit for translocation renal cell carcinoma.

  • Rhabdomyosarcoma

  • Among children younger than 20 years, the incidence of rhabdomyosarcoma is 4.3 new cases per 1 million population per year.

  • Rhabdomyosarcoma is the most common soft tissue sarcoma among children and adolescents.

  • Other benign and malignant soft tissue tumors must be excluded.

  • At pathologic examination, rhabdomyosarcoma must be differentiated from the other small, round, blue cell tumors of childhood (e.g., Ewing sarcoma, neuroblastoma, non-Hodgkin lymphoma).

  • Staging evaluation includes a thorough history and physical examination; CBC; imaging studies of the primary tumor, regional lymph nodes, lungs, and bones; and bone marrow examination.

  • Primary therapy is chemotherapy with surgery, radiation therapy, or both. Specific treatment depends on age of the patient, tumor primary site and histologic subtype, and extent of disease.

  • Mutilating surgery usually can be avoided because the tumor is sensitive to both chemotherapy and radiation therapy.

  • The OS rate exceeds 70%, but survival will depend on extent of disease. More than 90% of patients with localized, resectable, favorable histology tumors survive, but only about 25% of patients with metastatic disease survive.

  • Cure rarely is possible except for patients with botryoid tumors and those with embryonal tumors whose tumor arose in a favorable site and was completely resected at initial diagnosis.

  • Nonrhabdomyosarcoma Soft Tissue Sarcoma

  • Among persons younger than 20 years, 6.7 new cases occur per 1 million population per year.

  • Peaks in incidence occur among infants and among children older than 12 years of age.

  • Other tumors to be ruled out include benign soft tissue tumors, rhabdomyosarcoma, and extraosseous Ewing sarcoma.

  • Staging evaluation includes a thorough history and physical examination and imaging studies of the primary tumor and lungs.

  • Imaging of regional lymph nodes, liver, bones, and brain is indicated in some clinical settings.

  • Surgical excision with or without radiation therapy is indicated for patients with resectable tumors.

  • Chemotherapy may provide some benefit for patients with high-grade tumors larger than 5 cm in diameter and for those with unresectable or metastatic tumors.

  • Survival depends on the size and grade of the tumor, extent of disease, and extent of surgery that is possible. Patients with localized tumors 5 cm or less in diameter or localized low-grade tumors larger than 5 cm in diameter that can be resected have a survival rate exceeding 85%. Approximately 50% of patients with high-grade tumors larger than 5 cm or with unresectable disease survive. The survival rate is about 10% among patients with metastatic disease.

  • Cure generally is possible only for patients with local recurrence amenable to surgical extirpation and for those with distant recurrence of low-grade tumor that proves to be surgically resectable.

  • Retinoblastoma

  • Among children younger than 5 years of age, 11 new cases occur per 1 million population per year.

  • The two clinical forms of retinoblastoma are as follows: (1) hereditary, bilateral or multifocal (25% of cases)—this form, characterized by germline mutations of RB1 , may be inherited from an affected survivor or a silent carrier parent or may be the result of a new germline mutation, and (2) nonhereditary, unilateral or unifocal (75% of cases)—15% of unilateral cases represent germline mutations.

  • Clinical manifestations: Leukocoria is seen in more than 50% of cases and strabismus in 20% to 25%.

  • Differential diagnosis: Coats disease, retinopathy of prematurity, persistent hyperplastic primary vitreous, Toxocara uveitis, and toxoplasmosis need to be ruled out.

  • The approach to staging depends on tumor size and on the presence or absence of intraocular and extraocular extension.

  • Indirect ophthalmoscopic examination of both eyes with the patient under general anesthesia is essential.

  • Imaging studies helpful in staging include ultrasonography, orbital and cerebral CT, and MRI.

  • Bone marrow biopsy and cerebrospinal fluid examination are reserved for patients with extraocular disease, optic nerve involvement, or choroid invasion.

  • Treatment must be individualized and depends on laterality, potential for vision, and tumor extent.

  • Enucleation is reserved for cases in which no potential for useful vision remains. Cryotherapy and photocoagulation are useful for management of small primary or recurrent tumors.

  • Radiation therapy is the treatment modality of choice for controlling local disease and preserving vision with larger tumors.

  • Chemotherapy is used in the context of ocular-salvage treatments, with either systemic, intraarterial, or intravitreal administration, in patients with advanced intraocular disease with high-risk pathology, and in patients with extraocular disease.

  • Hepatoblastoma

  • Among children younger than 15 years of age, 2.6 new cases occur per 1 million population per year.

  • The mean age at diagnosis is 18 months.

  • The tumor is associated with familial adenomatous polyposis and Beckwith-Wiedemann syndrome.

  • An association with very low birth weight has been found.

  • The major histologic subtypes are epithelial or mixed, consisting of both epithelial and mesenchymal elements.

  • The epithelial subtype is further divided into fetal, embryonal, macrotrabecular, small-cell undifferentiated, and cholangioblastic.

  • Nearly all tumors show Wnt pathway activation, the majority having CTNNB1 -catenin ) mutations. Other recurrent somatic mutations occur in NFE2L2 and TERT.

  • An asymptomatic abdominal mass is present in most patients.

  • Other features may include anorexia, weight loss, vomiting, and precocious puberty (seen in 2% of cases).

  • Other malignant tumors to be excluded include hepatocellular carcinoma, embryonal sarcoma, rhabdomyosarcoma, angiosarcoma, and teratoma.

  • Benign tumors to be ruled out include hemangioma, hemangioendothelioma, hamartoma, and adenoma.

  • Staging evaluation includes a complete history and physical examination, with careful assessment for any congenital abnormalities or signs of precocious puberty.

  • Laboratory tests typically include CBC, serum chemical analysis, and α-fetoprotein and α-human chorionic gonadotropin assays.

  • Diagnostic imaging may include CT or MRI of the abdomen and chest CT.

  • Cure is possible only when complete surgical excision is performed. If complete excision is not feasible, liver transplantation should be considered.

  • Adjuvant chemotherapy (cisplatin based, in conjunction with doxorubicin or 5-fluorouracil) is useful preoperatively for achieving resectability and postoperatively for preventing distant metastasis.

  • Radiation therapy has a limited and ill-defined role.

  • Recurrent disease confers a poor prognosis, but repeated resection of local and metastatic recurrences has lengthened the survival period.

  • Hearing loss and nephrotoxicity (related to cisplatin), cardiac toxicity (related to doxorubicin), and development of second malignant neoplasms may occur.

  • With stage I/II disease , 4-year event-free survival (EFS) is 85% to 95%, and 4-year OS is 95% to 100%

  • With stage III disease , 4-year EFS and OS are approximately 60% to 70%.

  • With stage IV metastatic disease , 3- or 4-year EFS and OS are approximately 35% to 75% and 40% to 80%, respectively.

  • Adrenocortical Carcinoma

  • The incidence of adrenocortical carcinoma is two to three new cases per 10 million population per year in the United States.

  • A 10- to 15-fold higher incidence has been observed in southern Brazil.

  • It is associated with germline TP53 mutations and Li-Fraumeni syndrome.

  • Other conditions associated with pathologic androgen or cortisol production, such as Cushing syndrome or ovarian or testicular tumors, should be ruled out.

  • Evaluation for staging includes imaging studies of the abdomen, pelvis, and chest; skeletal scintigraphy; and measurement of blood and urine concentrations of adrenocortical hormones.

  • Complete surgical removal of the tumor is indicated.

  • Mitotane-based therapy or chemotherapy with cisplatin also is given.

  • Recurrent disease can be managed with further surgery, or experimental agents can be tried.

Solid tumors account for 60% of all pediatric malignant neoplasms, with approximately 3700 new cases diagnosed each year in the United States. The spectrum of tumor types that occur in children is much different from that observed in adults. Included among the diverse types of pediatric neoplasms are tumors of the central nervous system (CNS) (35%); neuroblastoma (15%); soft tissue sarcoma, including rhabdomyosarcoma (7%); Wilms tumor (6%); bone tumors, including osteosarcoma and Ewing sarcoma (8%); retinoblastoma (5%); and miscellaneous tumors, including hepatoblastoma, germ cell tumors, and melanoma (17%). Enormous progress has been made in the diagnosis and management of these tumors since the original demonstration of the chemosensitivity of Wilms tumor to actinomycin D in 1966. Cure rates for most childhood solid tumors have increased by as much as 50% since the mid-1970s. The increase is attributable largely to improved understanding of prognostically important biological and clinical features, enhanced precision of clinical staging systems, consistent use of supportive care, and development of more effective treatment, often incorporating a combination of chemotherapy, surgery, and radiation therapy.

Osteosarcoma

Epidemiology

Osteosarcoma, a malignant neoplasm derived from primitive mesenchymal cells and characterized by the presence of osteoid-producing spindle cell stroma, is the most common malignant bone tumor in the pediatric age group. Osteosarcoma ranks 10th among all newly reported pediatric cancers in the United States, accounting for close to 3% of all neoplasms in children and adolescents. The estimated annual incidence is 4.9 cases per 1 million population among white children and adolescents and 5.7 per 1 million population among African American children and adolescents. Most osteosarcomas occur during the first 2 decades of life, a period characterized by rapid skeletal growth. Boys are affected more commonly than girls. Several observations support the association between skeletal growth velocity and osteosarcoma. First, patients with osteosarcoma tend to be taller than their counterparts without this disease. Second, osteosarcoma develops at an earlier age in female patients than in male patients, perhaps because of differences in the timing of onset of puberty and the growth spurt.

Tumor Biology

Unlike osteosarcoma in adults, in whom more than 25% of tumors are associated with preexisting pathologic osseous conditions such as Paget disease or fibrous dysplasia, most pediatric osteosarcomas arise spontaneously in areas of bone without any abnormality. Irradiation is the best-characterized etiologic factor contributing to the development of secondary osteosarcoma. In a study involving 91 patients with second malignant bone sarcomas, osteosarcoma accounted for 72 cases, 52 (72%) of these tumors arising within previously irradiated fields. The median time for development of the secondary tumor was 9.6 years after irradiation. Osteosarcoma as a second malignancy is often associated with retinoblastoma; osteosarcoma is the most common malignancy in survivors of retinoblastoma, both in the irradiated and the nonirradiated areas, and it accounts for 25% to 40% of all second neoplasms in this population. One-half to two-thirds of osteosarcomas occur in the irradiated fields of the skull and face, one-third of tumors develop in the extremities, and fewer than 10% occur in the trunk. Osteosarcoma is also a very frequent malignancy in individuals with germline TP53 mutations (Li-Fraumeni syndrome)and in patients with REC helicase-associated disorders (Rothmund-Thomson, Werner, and Bloom syndromes).

Consistent with the association of osteosarcoma with retinoblastoma survivors and Li-Fraumeni syndromes, alterations in components of the cell-cycle control system appear to characterize the ontogeny of osteosarcoma. Studies using next-generation sequencing technologies show multiple somatic chromosomal lesions, including structural variations and copy number alterations. The single nucleotide variations exhibit a pattern of localized hypermutation (“kataegis”) in 50% of the tumors. Alterations in the p53 pathway are present in all tumors, and mutations in RB1 , ATRX , and DLG2 genes are noted in 29% to 53% of cases. The central role of those alterations in the pathogenesis of osteosarcoma is demonstrated in the animal models; genetically engineered mice based on osteoblast-restricted deletion of p53 and pRb develop short-latency high-grade osteosarcoma that reproduces many of the defining features of human osteosarcoma, including cytogenetic complexity and comparable gene expression signatures, histology, and metastatic behavior. Finally, pathway analysis and shRNA studies indicate that the PI3K/mammalian target of rapamycin (mTOR) pathway could represent a central vulnerability for therapeutic exploitation.

Pathology

Osteosarcoma is characterized by the presence of spindle cell stroma that produces osteoid. Conventional osteosarcoma can be subdivided histologically into three major groups depending on the predominant cell type. Approximately 50% of tumors are categorized as osteoblastic because the predominant extracellular element is osteoid, but 25% are chondroblastic , with a prominent cartilaginous component. Approximately 25% have a herringbone pattern similar to that observed in fibrosarcoma and are therefore called fibroblastic . No significant differences in overall outcome are apparent among these three histologic subtypes. In one report, however, patients with fibroblastic histology showed better histologic response and better overall outcome than were observed for patients with osteoblastic or chondroblastic histology.

Clinical Manifestations

Pain is the most common symptom in children and adolescents with osteosarcoma. Onset of pain often is insidious, and the pain usually involves the area affected by tumor. Severe pain of sudden onset commonly is associated with pathologic fracture. Swelling around the affected bone is the second most common clinical finding. The tumor may be easily palpable when located in areas such as the anterior surface of the femur but may manifest only as leg edema when occurring in difficult-to-appreciate areas such as the popliteal fossa. A painful limp that increases with weight bearing is the third most common symptom. Systemic signs and symptoms such as fever and weight loss are uncommon.

Osteosarcoma most commonly involves the long bones, most tumors occurring around the knee. The most frequent sites of involvement are the distal part of the femur, the proximal portion of the tibia, and the proximal part of the humerus. The axial skeleton, including the pelvis, is rarely affected in children (<10% of cases) but more frequently is involved in patients older than 60 years. Overt macroscopic metastatic disease occurs in 20% of cases and carries a grave prognosis.

Laboratory and Radiologic Evaluation

Laboratory evaluation often is unrevealing. Elevations of serum lactate dehydrogenase (LDH) and alkaline phosphatase levels are the most common laboratory abnormalities. The latter appears to correlate with osteoblastic activity and thus has proved useful in monitoring response to therapy.

Radiologic evaluation of a patient with osteosarcoma must include assessment of the primary site as well as a search for distant metastatic lesions. Plain radiography is the most effective method of detection of bone tumors. The main limitation of plain radiography is accurate delineation of local tumor extent. Characteristic radiologic findings in osteosarcoma commonly include a metaphyseal permeative lesion with periosteal new bone formation and destruction of preexisting cortical bone. A soft tissue mass is present in more than 90% of cases. Other radiologic signs commonly associated with osteosarcoma include cumulus cloudlike density and the presence of Codman triangle ( Fig. 92.1A ). A baseline chest radiograph should be obtained to search for distant metastatic lesions. Angiography usually is reserved for patients who receive intraarterial chemotherapy or for those who need optimal vessel visualization before limb salvage.

Figure 92.1, Osteosarcoma involving the right distal femur. (A) Plain radiograph shows a poorly defined, permeative destructive pattern, periosteal reaction, a Codman triangle, and an associated soft tissue mass. (B) Sagittal T2-weighted magnetic resonance image from the same patient shows abnormally dark area (representing increased signal intensity) in the intramedullary space consistent with tumor involvement by osteosarcoma. The associated posterior soft tissue mass is evident. Areas of bright focal signal enhancement represent intramedullary hemorrhage.

Computed tomography (CT) of the primary tumor is accurate in assessment of degree of tumor calcification and ossification, which are important in assessment of response to therapy. Chest CT always should be performed at the time of diagnosis for documentation of metastatic disease. Findings by magnetic resonance imaging (MRI) offer the best estimate of intramedullary tumor extension, joint and vascular involvement, detection of “skip” metastatic lesions, and delineation of the soft tissue component ( Fig. 92.1B ). The blood supply and vascularity of the tumor can be better appreciated with administration of gadopentetate dimeglumine, or gadolinium-diethylenetriamine pentaacetic acid (Gd-DTPA), a paramagnetic contrast material. MRI is helpful in assessing response to chemotherapy, as evidenced by changes in signal intensity on T 2 -weighted images and alterations in the enhancement of tumor tissue after administration of Gd-DTPA ( Fig. 92.2 ). Dynamic contrast-enhanced MRI (DEMRI) is a valuable method for assessing microcirculation in osteosarcoma. DEMRI can be used to evaluate changes in regional contrast access during chemotherapy. The findings appear to correlate accurately with histologic response and outcome, allowing early identification of patients at risk for recurrence. Radionuclide bone scans with technetium-99m ( 99m Tc)–labeled bone-seeking phosphate compounds are particularly useful in detection of metastatic bone lesions. Fluorodeoxyglucose–positron emission tomography (FDG-PET) is also a useful tool in diagnosis and response assessment.

Figure 92.2, Dynamic, paramagnetic, contrast-enhanced magnetic resonance image from a patient with an osteosarcoma of the distal right lower extremity. At subsequent pathologic examination, the focal areas of increased signal intensity were found to represent nests of viable tumor.

Biopsy of the primary tumor should be done carefully, preferably by the surgeon who will ultimately perform the definitive operation. In performance of biopsy of a suspected bone tumor, the following basic principles should be observed: (1) avoidance of transverse incisions, which can make subsequent surgery difficult; (2) avoidance of contamination of multiple compartments and hematoma formation because successful limb-sparing procedures can be jeopardized; and (3) if feasible, biopsy of the soft tissue component only.

Osteosarcoma Subtypes

In addition to conventional osteosarcoma, a small proportion of patients are seen with clinical subtypes that are characterized by distinct clinical, radiologic, and histologic characteristics.

Telangiectatic Osteosarcoma

Telangiectatic osteosarcoma is characterized microscopically by blood-filled spaces divided by septa containing neoplastic sarcomatous cells. Both radiologically and histologically, it is difficult to differentiate from aneurysmal bone cyst. Telangiectatic osteosarcoma accounts for fewer than 4% of cases of osteosarcoma. Age and anatomic distributions are similar to those in conventional osteosarcoma. On imaging studies, telangiectatic osteosarcoma manifests as a purely lytic lesion with a permeative destructive growth pattern; it usually disrupts the cortex but with minimal or no periosteal new bone formation, which often is multilayered, in an onionskin pattern. Overall, the radiographic appearance is not that of a typical osteosarcoma, which usually has a mixture of blastic and lytic areas. Indeed, a purely lytic radiographic appearance is a diagnostic requirement for its diagnosis. This pattern can simulate aneurysmal bone cyst. Histologically, telangiectatic osteosarcoma is very hemorrhagic, similar to the gross appearance of aneurysmal bone cysts. Microscopically, this tumor consists of cystlike spaces divided by septa, which are composed of highly atypical sarcomatous tissue. Unlike aneurysmal bone cysts, these cystic spaces have no endothelial lining, and the tumor cells are in direct contact with areas of hemorrhage. In the past, it was thought that telangiectatic osteosarcoma carried a worse prognosis than for conventional osteosarcoma. With appropriate multimodality therapy, however, the outcome is similar to or better than that for conventional osteosarcoma.

Low-Grade Intramedullary Osteosarcoma

Low-grade intramedullary osteosarcoma is a very rare variant of osteosarcoma, accounting for fewer than 1% of cases. Most cases are diagnosed during the third decade of life. Its anatomic distribution is similar to that of conventional osteosarcoma, with predilection for the distal femur and proximal tibia. In contrast with conventional osteosarcoma, symptoms typically develop over many months or even years before the patient comes to medical attention. Imaging studies usually show a variable pattern of lytic foci and dense, poorly demarcated areas. Periosteal new bone formation is minimal. Histologically, it shows a predominantly differentiated fibroblastic and osseous component, similar to what is seen in parosteal osteosarcoma (see “ Surface Osteosarcomas ” later). The differential diagnosis must distinguish this lesion from fibrous dysplasia. Treatment includes a complete resection of the lesion. Incomplete resection invariably results in local recurrence, and dedifferentiation increases with each recurrence.

Surface Osteosarcomas

Osteosarcomas that originate and grow predominantly on the surface of the bone include the parosteal, periosteal, and high-grade surface osteosarcomas.

Parosteal osteosarcoma is a low-grade tumor that grows predominantly on the surface of long bones, in an exophytic pattern. Because it is derived from the outer layer of the periosteum, it grows without causing elevation of the periosteum or evidence of periosteal new bone formation. Parosteal osteosarcomas account for 3% of all osteosarcomas. They tend to occur in skeletally mature patients, with diagnosis during the third and fourth decades of life. More than 80% of these tumors are located in the distal portion of the femoral shaft, in its posterior aspect, within the superior popliteal area. Because of their slow growth, parosteal osteosarcomas usually manifest as a painless mass. Imaging shows a tumor growing on the surface of the bone, with a broad base, in a mushroom-like fashion. The mass typically is densely mineralized and has lobulated outlines. Microscopically, these tumors are characterized by the presence of a spindle cell fibroblastic component, with variable osteoid production, low mitotic rate, and no atypical features. The treatment is surgical, and a complete resection is mandatory. Incomplete resections invariably lead to local recurrences, and the risk of high-grade transformation increases with local recurrence.

Periosteal osteosarcoma is a low- to intermediate-grade tumor that has a predominantly chondroblastic differentiation. It originates in the deep layer of the periosteum, so its growth is manifested by a separation and elevation of the periosteum from the cortex, causing a prominent periosteal new bone formation. It represents fewer than 2% of all osteosarcomas. Periosteal osteosarcoma is a tumor of childhood, with a peak incidence during the second decade of life, and it has a female predominance. The usual location is in the long bones of the lower extremity, most commonly the tibia, with an affinity for the diaphysis. The presentation is similar to that with conventional osteosarcoma, with pain and swelling of short duration. Imaging studies show a predominantly fusiform lesion on the surface of a long bone. Treatment is usually the same as for conventional osteosarcoma.

High-grade surface osteosarcoma is a very rare subtype of surface osteosarcoma, accounting for fewer than 1% of all osteosarcomas. The age distribution, anatomic location, and clinical presentation are similar to conventional osteosarcoma. By imaging, high-grade surface osteosarcoma simulates a periosteal osteosarcoma. Microscopically, however, it is indistinguishable from conventional osteosarcoma, showing anaplastic cellular features and osteoid and immature bone and cartilage formation. Medullary extension is minimal. Treatment should be the same as for conventional osteosarcoma.

Prognostic Factors

The most important adverse prognostic factor in patients with osteosarcoma is the presence of metastatic disease. In addition, primary tumor location is associated with outcome. Children with primary tumors of the tibia and distal femur appear to have a more favorable prognosis than those with axial primary tumors. This finding highlights the importance of complete surgical resection in the management of this malignant disease. For patients with localized disease, factors associated with poor prognosis include measures of tumor burden, such as tumor size, and levels of alkaline phosphatase and LDH, as well as more biological measures, such as poor histologic response to preoperative chemotherapy, hyperdiploidy, and increased expression of P-glycoprotein or Ki-67. The percentage of tumor necrosis after preoperative chemotherapy is the most consistent and important factor associated with outcome in children and adolescents with localized osteosarcoma. A favorable response (>90% tumor necrosis) correlates with excellent overall survival (OS) rate. Patients who have less than 90% tumor necrosis are considered poor responders, for whom the prognosis usually is poor. Because of this strong correlation between degree of histologic response to preoperative chemotherapy and outcome, noninvasive methods such as DEMRI, used for monitoring tumor response, can be used to evaluate changes in regional contrast access during chemotherapy. Finally, a proportion of patients with extremity osteosarcoma are seen with a pathologic fracture, or a pathologic fracture develops after institution of therapy. This has been considered a poor prognostic factor and an indication for immediate amputation. It is possible, however, that with the use of preoperative chemotherapy and judicious use of limb-sparing techniques, a selected group of patients with pathologic fracture may still do well without amputation.

Treatment

Optimal management of osteosarcoma consists of multiple-agent chemotherapy and local control measures, including amputation or limb-sparing surgical procedures. Box 92.1 outlines the standard approach to management of osteosarcoma used at most institutions. Before the development of limb-sparing procedures, amputation was the standard surgical method used for curative treatment of osteosarcoma. Amputation now generally is reserved for primary tumors deemed unresectable. Limb function after below-the-knee amputation usually is excellent. Over the past several years, the role of limb-sparing procedures has increased dramatically. As a result of refinements in neoadjuvant chemotherapy, bioengineering, and imaging techniques, it is estimated that as many as 80% of patients with osteosarcoma will eventually be candidates for limb-sparing procedures. The criteria for limb-sparing procedures include (1) absence of major neurovascular involvement by tumor, (2) feasibility of wide surgical excision to include a normal muscle cuff in all directions and en bloc removal of all biopsy sites, (3) resection of the adjacent joint and capsule, (4) adequate motor reconstruction with regional muscle transfer, and (5) adequate soft tissue coverage. In the past, immature skeletal age and primary tumor of the humerus were relative contraindications; however, new expandable prosthetic devices may help overcome this problem. More recent improvements have been concentrated on achieving noninvasive extension of prostheses. One such method is the Phenix technology (Repiphysis). The basic principle involves storage of energy in a spring maintained in compressed form by a locking system. Prosthetic lengthening is performed through exposure to an external electromagnetic field that pilots the locking system and allows controlled release of the spring energy ( Fig. 92.3 ). Prosthetic expansion of several millimeters can be achieved with each procedure, and the total duration of the procedure is less than 30 seconds ( Fig. 92.4 ). Thus limb-sparing procedures are considered feasible in the care of most children and adolescents with osteosarcoma; when these procedures are appropriately performed, the risk of local recurrence is low (<5%). Long-term functional outcome, however, must be carefully compared with that obtained with amputation alone. Complications of limb-sparing surgery include infection, nonunion, fracture, and unstable joints.

Box 92.1
St. Jude Children's Research Hospital Approach to Management of Osteosarcoma

At St. Jude Children's Research Hospital, the standard approach to the management of osteosarcoma entails the use of a platinum-based chemotherapy regimen and radical resection of the primary tumor. The chemotherapy regimen used consists of high-dose methotrexate, cisplatin, and doxorubicin (MAP). Resection for local control is performed after completion of three courses of chemotherapy, and postoperative chemotherapy is not modified on the basis of histologic response unless progression of the disease is documented. Most patients with extremity osteosarcoma are candidates for limb salvage surgery. Amputation is performed only in selected cases.

In patients with metastatic disease at diagnosis, aggressive resection of the primary tumor and all sites of metastasis is indicated. For patients with unresectable disease, radiation therapy is used, although the outcome for this group of patients is very poor.

The few patients with local recurrence after a limb-sparing procedure undergo amputation. Recurrent disease to the lungs is managed with aggressive surgery, and repeated thoracotomy usually is needed. Systemic chemotherapy with second-line regimens is added in cases of early recurrence or unresectable disease.

Figure 92.3, Activation of Phenix (Repiphysis) device expansion in treatment for osteosarcoma involving the left lower extremity in a 13-year-old boy. The activating device is on the left. The electromagnetic coil is being held over the annular protuberance within the implant.

Figure 92.4, Radiographs show the expandable portion of the Phenix (Repiphysis) device. The titanium tubular portion expands out of the polymeric tube. The increased distance between barrel and stem indicates the amount of expansion that occurred (arrows).

Before the introduction of adjuvant chemotherapy, fatal metastatic disease developed in more than 80% of patients with osteosarcoma. Trials of single-agent chemotherapy began in the 1960s and early 1970s and established, in a nonrandomized manner, a role for the use of chemotherapy in the management of osteosarcoma. Responses with single-agent high-dose methotrexate or doxorubicin occurred in 20% to 40% of patients with metastatic disease. Since then, different combinations of platinum compounds, doxorubicin, and high-dose methotrexate have formed the basis of standard chemotherapy regimens that lead to cure in 50% to 75% of patients with nonmetastatic disease ( Table 92.1 ).

Table 92.1
Summary of Most Relevant Chemotherapy Protocols for the Management of Osteosarcoma
T reatment
Protocol/Author N Preoperative Postoperative a Outcome Comment(s)
T-10/Meyers et al (1992) 31 MTX-BCD GR : + DOX
PR : + CDDP/DOX
5-yr EFS: 73% PR : No advantage of intensified postoperative treatment
T-12/Meyers et al (1998) 36 MTX-BCD-DOX-CDDP Same 5-yr EFS: 78% No advantage of intensified neoadjuvant treatment
MIOS/Link et al (1986) 77 None MTX-BCD-DOX-CDDP 2-yr DFS: 66% Randomized study demonstrating need for chemotherapy
36 Surgery alone 2-yr DFS: 17%
CCG-782/Provisor et al (1997) 268 MTX-BCD GR : + DOX
PR : + DOX-CDDP
8-yr EFS: 53%
GR : 81%
PR : 46%
PR : No advantage of intensified postoperative treatment
COSS-82 arm A/Winkler et al (1988) 59 MTX-BCD GR : Same
PR : + CDDP/DOX
4-yr EFS: 49%
COSS-82 arm B/Winkler et al (1988) 66 MTX-DOX-CDDP GR : Same
PR : + IFO
4-yr EFS: 68% PR : IFO not advantageous
COSS-86 low risk/Fuchs et al (1998) 41 MTX-DOX-CDDP Same 10-yr EFS: 66% No difference for IA vs IV administration of CDDP
COSS-86 high risk/Fuchs et al (1998) 128 MTX-DOX-CDDP-IFO Same 10-yr EFS: 67% Use of IFO for high-risk patients
EOI-1/Bramwell et al (1992) 142 CDDP-DOX × 3 CDDP-DOX × 3 5-yr DFS: 57% Importance of a two-drug short regimen with CDDP-DOX
140 CDDP-DOX-MTX × 2 CDDP-DOX-MTX × 2 5-yr DFS: 41%
EOI-2/Souhami et al (1997) 192 MTX-DOX + BCD-CDDP 5-yr PFS: 44% Two-drug short regimen may be better than longer; more complex protocols
199 CDDP-DOX Same 5-yr PFS: 44%
IOR-1/Ferrari et al (1997) 127 MTX-CDDP GR : + DOX-BCD
PR : MTX-DOX-BCD
12-yr DFS: 46% HD MTX better than MD MTX
IOR-2/Bacci et al (2000) 164 MTX-CDDP-DOX GR : same 5-yr DFS: 63% Importance of dose intensity
PR : + IFO/ETO GR : 67% PR : IFO/ETO good salvage
PR : 56%
IOR-3/Bacci et al (1997) 139 MTX-CDDP-DOX GR : same PR : + IFO 3-yr DFS: 60% PR : No benefit of IFO
IOR-4/Bacci et al (2000) 133 MTX-CDDP-DOX-IFO Same 5-yr EFS: 56% No benefit of intensified therapy with IFO
SJ-OS91/Meyer (2001) 47 CBP-IFO Same + DOX-MTX 3-yr EFS: 72% CBP is a good alternative to CDDP for nonmetastatic disease
INT 0133/Meyers (2005) 577 Randomized to: 5-yr. EFS: Possible synergistic effect between IFO and MTP-PE
A : MTX-CDDP-DOX 64%
A +: MTX-CDDP-DOX + L-MTP-PE 63%
B : MTX-CDDP-DOX + IFO 56%
B +: MTX-CDDP-DOX + IFO + L-MTP-PE 72%
SSG VIII/Smeland (2003) 113 MTX-CDDP-DOX GR : same 5-yr EFS: 68%
PR : + IFO-ETO 5-yr EFS: 53% PR : Lack of benefit of modifying postoperative therapy
SJ-OS99/Daw et al (2011) 72 CBP-IFO-DOX Same 5-yr EFS 66.7% Treatment without HD-MTX and CDDP
EURAMOS-1/Bielack et al (2015) ; Marina et al (2016) 2260 MTX-CDDP-DOX GR: +/- IFN alpha-2b 3-yr EFS: 75% (no IFN) vs. 77 (+ IFN) Lack of benefit of addition of IFN alpha-2b for good responders
PR: +/- IE Lack of benefit of addition of IE to poor responders.
BCD, Bleomycin, cyclophosphamide, dactinomycin; CBP, carboplatin; CCG, Children's Cancer Group; CDDP, cisplatin; COSS, Cooperative Osteosarcoma Study; DFS, disease-free survival; DOX, doxorubicin; EFS, event-free survival; EOI, European Osteosarcoma Intergroup; ETO, etoposide; GR, good histologic responders; HD, high-dose; IA, intraarterial; IE, ifosfamide and etoposide; IFN, interferon; IFO, ifosfamide; INT, Intergroup; IOR, Istituto Ortopedico Rizzoli; IV, intravenous; L-MTP-PE, liposomal-encapsulated muramyl tripeptide phosphatidylethanolamine; MD, moderate dose; MIOS, Multi-Institutional Osteosarcoma Study; MTX, methotrexate; N, number of patients; PFS, progression-free survival; PR, poor histologic responders; SJ-OS, St. Jude Osteosarcoma; SSG, Scandinavian Sarcoma Group.

a Postoperative regimens are either the same as or in addition to preoperative regimens or as specified.

For more than 2 decades, therapy for nonmetastatic osteosarcoma has followed the basic guidelines of the T-10 protocol, and most of the current treatment strategies have evolved from lessons learned with it. The T-10 protocol and its variants consist of a multiple-agent regimen of high-dose methotrexate, doxorubicin, cisplatin, and a combination of bleomycin, cyclophosphamide, and dactinomycin. When the T-10 protocol guidelines are used, the 5-year actuarial event-free survival (EFS) rate may approach 70%. These results are not always reproducible, however, and multiinstitutional US and European studies conducted according to similar guidelines have shown lower EFS rates, usually 50% to 60%. The lack of reproducibility of the results of the original study may be attributed in part to the complexity of the treatment. In a study performed by the European Osteosarcoma Intergroup, patients were randomly assigned to receive the T-10 protocol or a much simpler treatment of 6 courses of a combination of cisplatin and doxorubicin. Only half of the patients in the T-10 protocol–like group completed all scheduled therapy, but 94% of the patients in the short-treatment group received the 6 courses of chemotherapy. The results for both treatment groups were identical, suggesting that a simple, two-drug regimen of cisplatin and doxorubicin can cure more than half of patients with nonmetastatic osteosarcoma.

A major contribution of the T-10 protocol and its predecessor T-7 is that the histologic response to neoadjuvant chemotherapy has been identified as the most important prognostic factor in the care of patients with nonmetastatic disease. Intensification of postoperative or preoperative chemotherapy with cisplatin, doxorubicin, and ifosfamide, however, has not improved outcome. To increase the proportion of good histologic responders, some researchers have investigated intraarterial administration of cisplatin. In the context of an aggressive, multiple-agent treatment, however, intraarterial infusion of cisplatin does not offer any significant advantages.

In recent years, ifosfamide has been incorporated into therapeutic regimens for osteosarcoma. After early reports showed that ifosfamide as a single agent achieved response rates of 10% to 60% in patients with advanced or treatment-refractory osteosarcoma, some investigators began to use the agent in salvage therapy regimens for poor responders. More recently, ifosfamide has been incorporated into frontline therapy in many regimens. However, whether incorporating the agent into standard treatment will improve the cure rate is unknown.

In the cooperative POG-Children's Cancer Group (CCG) Intergroup INT0133 trial, patients with localized osteosarcoma received treatment with a standard methotrexate–cisplatin–doxorubicin regimen and underwent a double randomization. The first randomization was to receive ifosfamide; the second randomization was to receive MTP (muramyl tripeptide), a derivative of the Bacillus Calmette-Guérin (BCG) cell wall that is known to stimulate the immune system, with the objective of decreasing pulmonary failures. Therefore four regimens were compared. The addition of ifosfamide or MTP alone to the standard regimen did not seem to provide any benefit. However, the addition of both ifosfamide and MTP-PE (muramyl tripeptide phosphatidylethanolamine) resulted in a significantly better EFS compared with the other three regimens. When the long-term outcome of the same cohort of patients was analyzed, patients who received MTP had improved OS (although not improved EFS) compared with those who did not. The significance of these findings is unclear.

The combination of ifosfamide and etoposide appears to be more effective than ifosfamide alone in patients with advanced or refractory osteosarcoma. In patients with refractory osteosarcoma, some of whom had previously received ifosfamide, the combination of ifosfamide and etoposide achieved response rates of 15% to 48%. Response rates of 59% have also been reported in patients with untreated metastatic osteosarcoma receiving this combination.

These results have led to the increasing use of this combination for the salvage of patients with poor histologic responses; however, evidence is lacking for improved outcome in this group of patients. The recently reported European-American Osteosarcoma (EURAMOS) trial for patients with localized osteosarcoma used a MAP (high-dose methotrexate, Adriamycin, cisplatin) backbone and randomized patients with good response (<10% viable tumor after 2 MAP cycles) to the addition of interferon alpha 2b, and patients with poor response (≥10% viable tumor) to the addition of ifosfamide and etoposide. Neither intervention resulted in improved outcomes over the standard MAP; the 3-year EFS estimates for good responders were 75% and 77% for MAP and MAP plus interferon alpha 2b, respectively, and the 3-year EFS estimates for poor responders were 55% and 53% for MAP and MAP plus ifosfamide and etoposide, respectively.

Methotrexate is the classical antifolate and blocks the action of dihydrofolate reductase. Methotrexate was among the first drugs reported to have an antitumor effect in osteosarcoma, and it has been a major component of the osteosarcoma treatment ever since. Furthermore, there is evidence that suggests that the pharmacokinetics of this agent may influence outcome, although this influence is minor in the context of more intensive multiagent protocols. However, the role of methotrexate appears to be limited to the administration of high doses (usually 12 g/m 2 ) because the administration of lower doses appears to have a lesser therapeutic effect. The administration of high-dose methotrexate requires close monitoring of serum levels, which cannot be performed in many institutions, particularly those of developing countries, and extensive supportive measures, including hyperhydration, urine alkalinization, and leucovorin rescue, which must be adjusted to the methotrexate serum levels. The development of acute renal failure, mediated by the perception of methotrexate and its metabolites in the renal tubules, is a potentially life-threatening complication. Even among patients receiving all available monitoring and supportive care, the rate of severe nephrotoxicity is 2%, and the mortality rate is 4.4%. Dialysis methods have limited effectiveness in removing methotrexate compared with the rapid reductions in plasma concentrations that can be achieved with carboxypeptidase G2. Studies have shown the possibility of obtaining comparable outcomes without the use of methotrexate, provided intensification of platinum, ifosfamide, and doxorubicin is performed.

Cisplatin is one of the most active agents against osteosarcoma. The toxicity of this agent is substantial and includes hearing loss and renal impairment that can be permanent for some patients. Substitution of carboplatin for cisplatin was investigated at St. Jude Children's Research Hospital. In the context of a multiple-agent chemotherapeutic approach with high-dose methotrexate, ifosfamide, and doxorubicin, use of carboplatin resulted in a 3-year EFS rate of 72%, an outcome comparable with that obtained with cisplatin-based therapy but with less long-term toxicity. Incorporation of carboplatin in future osteosarcoma trials needs further evaluation, however, because in some reports, carboplatin as a single agent has been found to have poor antitumor effect in the treatment of metastatic disease.

Approximately 20% of patients with osteosarcoma have clinically detectable metastatic disease at diagnosis, and their outcome usually is very poor. Optimal treatment for these patients entails a very aggressive multimodality approach that combines intensive preoperative and postoperative chemotherapy with resection of both the primary tumor and metastatic lesions. When these guidelines are followed, contemporary protocols that incorporate ifosfamide or the combination of ifosfamide and etoposide, along with high-dose methotrexate, doxorubicin, and cisplatin, result in 2- to 5-year progression-free survival rates of 25% to 45%.

For patients with recurrent osteosarcoma, a very aggressive surgical approach is recommended; the 5-year postrelapse survival rate for patients in whom a complete resection of all macroscopic disease can be achieved is close to 40%. In a small number of patients (3% to 5%), metachronous osteosarcoma can develop after primary treatment. This condition has been associated with the presence of germline TP53 (Li-Fraumeni) or RB1 gene mutations. These second primaries usually develop within 3 years from initial diagnosis, and with systemic chemotherapy and surgery, 30% to 40% of these patients can be cured.

Lung metastasis develops in most patients in whom therapy fails. The ability to control pulmonary micrometastatic disease after completion of therapy would certainly result in a significant improvement in outcome. In an animal model, administration of liposome-encapsulated muramyl tripeptide phosphatidylethanolamine (L-MTP-PE) resulted in activation of pulmonary macrophages and eradication of pulmonary micrometastatic lesions. However, as indicated earlier, the clinical impact of the addition of this agent in the frontline of osteosarcoma therapy is not well defined. Other strategies under evaluation include administration of aerosolized granulocyte-macrophage colony-stimulating factor, although available data seem to indicate lack of effect of such approach.

Finally, administration of high doses of the bone-seeking radio-pharmaceutical samarium-153 ethylenediamine tetramethylene phosphonate ( 153 Sm-EDTMP) may provide good pain palliation with minimal nonhematologic toxicity for patients with local recurrences or bone metastasis of osteosarcoma, and its role may be expanding.

Ewing Sarcoma Family Tumors

Epidemiology

The term Ewing sarcoma family tumors (ESFTs) defines a group of small, round cell neoplasms of neuroectodermal origin that manifest as a continuum of neurogenic differentiation. On this continuum, Ewing sarcoma of bone represents the least differentiated (most primitive) form of neuroectodermal tumors, and peripheral neuroepithelioma represents the most differentiated form. Ewing sarcoma is the second most common malignant bone tumor in children and adolescents. The estimated incidence among white children and adolescents (0 to 19 years old) is 2.95 cases per 1 million population. The tumor is rare in the nonwhite population; its incidence is 3.44 per million white children and adolescents compared with 0.65 in African Americans, and it also shows a lower incidence in Hispanics compared with non-Hispanics (2.60 versus 3.03, respectively). Males are predominantly affected in most series; most cases are diagnosed during the second decade of life. The location of the tumor varies. Although most Ewing sarcomas arise in bone, a significant proportion arise in soft tissue. The most common locations for this tumor are the chest wall, pelvis, and extremities, but any bone can be involved.

Tumor Biology

The histogenesis of Ewing sarcoma has been a source of controversy since the first description of the tumor in 1921. Various hypotheses have been proposed in an attempt to identify the possible cell of origin in Ewing sarcoma. Among these, cells of endothelial, pericytic, myeloid, mesenchymal, and neuroectodermal origin have been suggested. The existence of either a mesenchymal stem cell or an early primitive neuroectodermal cell that has retained its ability for multilineage differentiation is the currently accepted hypothesis. It is now well accepted that ESFTs constitute a single group of neurally derived neoplasms that share unique immunocytochemical, cytogenetic, and molecular markers.

Nearly all ESFTs have a reciprocal translocation that involves the EWS gene in chromosome 22q12. The t(11;22) is the most commonly observed translocation (85% to 95% of cases) and juxtaposes the DNA-binding domain of the human homologue of the murine Fli1 gene in chromosome 11 with the 5′ end of EWS in chromosome 22 ( Fig. 92.5 ). The most common fusions occur between exon 7 of EWS and exon 6 of FLI1 (type 1; 55% to 60% of cases) and between exon 7 of EWS and exon 5 of FLI1 (type 2; 25% of cases). The t(11;22) appears to play a pivotal role in development of ESFTs because the EWS-FLI1 fusion transcript can transform NIH 3T3 cells. As many as 10% of ESFTs contain an alternative translocation between chromosomes 21 and 22. This translocation fuses the ERG gene on chromosome 21 and the EWS gene on chromosome 22, producing an ERG-EWS fusion transcript. Finally, rare cases of ESFT contain a t(7;22)(p22;q12) that fuses the ETV1 and EWS genes. The mechanism by which these fusion transcripts become tumorigenic is poorly understood, although it has been postulated that this chimeric transcript can transcriptionally deregulate members of the manic fringe family of genes, which are instrumental in somatic development. The IGF1/IGF1R pathway is actively involved in the cell transformation and inhibition of apoptosis induced by EWS-FLI1 . Next-generation sequencing studies have reported a more comprehensive description of the genomic landscape of these tumors. In these studies, ESFT has demonstrated to have low number of single-nucleotide and structural variants. However, chromosome 1q gain and chromosome 16q loss appear to be associated with worse prognosis, and single nucleotide variants and structural variants correlate with age at diagnosis and survival. The most frequent coding variants occur in STAG2 , TP53 , and in epigenetic regulators, and coexistence of STAG2 and TP53 mutations is associated with highly aggressive tumors.

Figure 92.5, Karyotype obtained from a Ewing sarcoma tissue specimen shows the characteristic t(11;22) (q24;q12), secondary t(1;16)(q21;q13), and trisomy 8.

Pathology

Microscopic examination shows that Ewing sarcoma is the prototypical small, round, blue cell tumor of childhood. Ewing sarcoma is characterized by the presence of a dimorphic pattern of densely packed cells with variable amounts of large clear cytoplasm. Individual cells have an ellipsoid nucleus without distinct cytoplasmic outlines. The cells are primitive, show a paucity of organelles, and often contain large amounts of intracellular glycogen. Various microscopic patterns, including the diffuse, lobular, organoid, and filigree patterns, have been described. Immunocytochemical analysis shows vimentin and CD99 reactivity. The latter monoclonal antibody specifically recognizes the cell surface antigen p30/p32 MIC-2, which normally is expressed as a component of the T-cell receptor complex. When overt neural differentiation is present, the term peripheral neuroectodermal tumor or peripheral neuroepithelioma is used. In such cases, the cells are round with more abundant cytoplasm but do not show mature neural elements, such as nerve bundles and mats of neuropile. The term extraosseous Ewing sarcoma has been reserved for neoplasms that cannot be differentiated from Ewing sarcoma at light microscopic examination but arise exclusively in soft tissue. Patients with this histologic variety accounted for 5% of all subjects in the Intergroup Rhabdomyosarcoma Study, although extraosseous tumors are now managed in the same manner as for classic Ewing sarcoma of bone. Despite these histologic distinctions, Ewing sarcoma of bone, extraosseous Ewing sarcoma, primitive neuroectodermal tumors, and peripheral neuroepithelioma are unified by the presence of the same oncogenic events and should therefore be considered the same neoplasm and managed in a similar manner. The advent of molecular diagnostic techniques has allowed development of sensitive tests, such as reverse transcriptase-polymerase chain reaction (RT-PCR) analysis, which is accurate in detection of EWS-FLI1 and EWS-ERG fusion transcripts at levels well below those commonly found with routine cytogenetic studies. RT-PCR assay is a useful adjunct in the diagnosis of Ewing sarcoma, particularly in differentiation from other soft tissue small, round cell tumors, such as rhabdomyosarcoma.

Clinical Manifestations

Ewing sarcoma commonly manifests during the second decade of life (median age, 13 years) with localized pain and a visible palpable mass. Fewer than 3% of cases occur in children younger than age 3 years. Boys are more commonly affected than girls. Pathologic fractures may be present in as many as 15% of children and adolescents before diagnosis. Back pain, extremity weakness, or altered sensation should raise suspicion for the presence of primary or metastatic disease. Systemic manifestations such as fever are more frequent than in osteosarcoma. Almost half of patients have signs and symptoms referable to the primary tumor for more than 3 months before the diagnosis is made. Ewing sarcoma has a tendency to involve the shaft of long tubular bones, pelvis, and ribs, but almost every bone can be affected. More than 50% of the tumors arise from axial bones, the pelvis being the most commonly involved (25%); one-third of the tumors originate in the lower extremities and fewer than 10% in the upper extremities. Chest wall Ewing sarcoma is known as Askin tumor ( Fig. 92.6 ). Approximately 20% to 25% of patients with this tumor have metastatic disease when they are seen for evaluation. The most common sites of metastatic disease are in the lungs followed by the bones and bone marrow. Metastatic disease appears to be associated with older age and with large tumor or pelvic primaries.

Figure 92.6, Primitive neuroectodermal tumor involving the chest wall. (A) Plain radiograph shows complete opacification of the right hemithorax, marked mediastinal shift, and slight tracheal narrowing. (B) Coronal T1-weighted magnetic resonance image shows a large right intrathoracic mass with subjacent pleural effusion and small localized area of normal aerated lung parenchyma.

Laboratory and Radiologic Evaluation

Patients with suspected Ewing sarcoma should be thoroughly evaluated to define the extent of local disease and the presence of metastatic lesions. Initial laboratory studies include complete blood cell count and erythrocyte sedimentation rate; measurement of serum electrolytes; LDH assay; renal and liver function tests; determination of alkaline phosphatase, calcium, phosphorus, and magnesium levels; and coagulation profile. In Ewing sarcoma, elevation of erythrocyte sedimentation rate and of serum LDH is not uncommon. Bone marrow aspiration and biopsy should be performed, and evaluation with molecular techniques such as RT-PCR is recommended. Important imaging studies are chest radiography, plain radiography of primary and metastatic sites, bone scintigraphy, CT of the chest, and MRI of the primary site with T 1 - and T 2 -weighted sequences, as well as DEMRI.

In Ewing sarcoma, plain radiographs typically show a diaphyseal destructive lesion with a laminated periosteal reaction and large soft tissue mass ( Fig. 92.7A ). St. Jude Children's Research Hospital favors MRI over CT for defining the intramedullary component of the primary tumor and the extent of soft tissue mass ( Fig. 92.7B ). In contrast with the situation in osteosarcoma, DEMRI findings are not a reliable prognostic indicator. Newer techniques such as positron emission tomography (PET) appear to be useful in noninvasive evaluation of response to chemotherapy.

Figure 92.7, Ewing sarcoma of the right distal tibia. (A) Plain radiograph shows a poorly marginated destructive lesion with associated periosteal reaction. (B) Sagittal T1-weighted magnetic resonance image of the same lesion shows a well-demarcated dark area of low signal intensity defining the intramedullary extent of the tumor contrasted against the bright appearance of the high signal intensity derived from intramedullary fat. The soft tissue component of the tumor is evident in the posterior aspect.

Prognostic Factors

A variety of initial clinical features have been correlated with outcome of therapy for Ewing sarcoma. Several groups have analyzed factors that are prognostically important in Ewing sarcoma. The results confirmed the classic clinical features associated with poor prognosis, such as metastatic disease, older age, large tumor size, and trunk and pelvic primary sites. However, with refinement in the multidisciplinary approach to this disease that entails newer and more intensive chemotherapeutic regimens and superior local control measures, some of these classic prognostic factors are being redefined.

Although large tumor size classically has been associated with worse prognosis, this feature may be less important with more aggressive treatment. For example, in the first Cooperative Ewing Sarcoma Study (CESS-81), tumors larger than 100 cm 3 were associated with worse outcome. With treatment improvements, the tumor size associated with worse prognosis has increased to 200 cm.

Tumor location also seems to be losing its prognostic significance with newer treatments, although pelvic and axial tumor locations classically were associated with worse outcome in the early studies. differences in outcome were minimal in subsequent studies. The first POG–CCG Ewing sarcoma trial (POG-8850/CCG-7881), an investigation of the effect of addition of ifosfamide and etoposide to the standard VACD regimen (vincristine, actinomycin D, cyclophosphamide, and doxorubicin), showed that addition of the drug pair abrogated the negative prognostic implications of large tumor size (>8 cm in diameter) and pelvic location. Of note, the benefit of addition of ifosfamide and etoposide was not seen in patients older than 18 years.

A newer prognostic factor is degree of histologic response to chemotherapy. European studies consistently have shown the prognostic value of histologic response, the significance of which stands across protocols and appears to be independent of the drugs used. Patients with good histologic responses had a significantly better outcome than those with poor responses in the consecutive REN-1, -2, and -3 trials in Italy and in the CESS-81 and CESS-86 trials in Germany. Because ESFTs manifest a continuum of neuroectodermal differentiation, the histologic diversity may reflect different biological behaviors. Evidence is lacking, however, for correlation of degree of neuroectodermal differentiation with prognosis. The most important prognostic factor remains the presence of metastatic disease at diagnosis. Advances in management of ESFTs have resulted in only very modest improvement in outcome among patients with metastatic lesions. Even among patients with metastatic disease, heterogeneity is common. With an appropriately intensive treatment that includes bilateral lung irradiation, the European Intergroup Cooperative Ewing Sarcoma Studies (EICESSs) have shown that patients with isolated metastatic lesions in the lung may have a better prognosis, albeit still inferior to that for patients with localized disease. Patients with extrapulmonary metastasis have a worse prognosis.

With the use of molecular techniques in the staging of ESFTs, it is evident that a significant proportion of patients with localized disease (20% to 30%) have micrometastatic disease in the bone marrow detected by polymerase chain reaction (PCR) assay. In patients with lung metastasis and those with bone metastasis, this proportion is 40% and 90%, respectively. Although clinical stage remains the most powerful prognostic indicator in patients with ESFT, detection of minimal disease in the bone marrow by RT-PCR has been associated with metastatic disease and appears to connote a more adverse clinical outcome in a subset of patients with apparently localized disease. The utility of detecting circulating neoplastic cells by RT-PCR in patients with ESFT is less clear.

Treatment

Before the introduction of systemic chemotherapy, fewer than 20% of children with Ewing sarcoma treated with either surgery or radiation therapy alone were expected to be long-term survivors. In the past 3 decades, major advances have been made in the management of Ewing sarcoma. These advances derive largely from cooperative trials ( Table 92.2 ). The first studies conducted by the Intergroup Ewing Sarcoma Study (IESS) showed the importance of adjuvant chemotherapy that included a combination of alkylating agents and anthracyclines (IESS-I and IESS-II). Thereafter, conventional management of Ewing sarcoma included local control measures in combination with administration of VACD. More recently, incorporation of ifosfamide (VAID [vincristine, actinomycin D, ifosfamide, doxorubicin]) has resulted in modest benefit for patients with high-risk features. Combined administration of ifosfamide and etoposide (IE) has been shown to be very active in patients with Ewing sarcoma who have not received previous treatment. Two randomized studies were conducted to investigate the effect of adding etoposide to VACD or VAID regimens. In EICESS-92, patients with localized high-risk disease (tumor diameter >200 cm 3 ) receiving VAID did not seem to benefit from the addition of etoposide. The first POG–CCG Ewing sarcoma trial (POG-8850/CCG-7881) was an investigation of incorporation of the IE combination in frontline management of ESFTs. Patients were randomly assigned to receive VACD with or without IE; patients receiving IE plus VACD appeared to have a more favorable outcome.

Table 92.2
Management of Nonmetastatic Ewing Sarcoma Family of Tumors
Study Patients ( n ) Regimen Results: 5-Year DFS Comment(s)
IESSS
IESS-I (1973–1978) 342 VAC 24% Value of doxorubicin
VAC + WLI 44% Worse results for pelvic ESFT
VACD 60% Benefit of WLI?
IESS-II (1978–1982) 214 VACD-HD 73% Value of aggressive cytoreduction
VACD-MD 56%
First POG-CCG (1988–1993) NA VACD 54% Value of combination IE
VACD + IE 69%
Second POG-CCG (1995–1998) 492 VCD + IE 48 wk 75% No differences between standard and dose-intensified therapy
VCD + IE 30 wk 76%
S t . J ude S tudies
ES-79 (1978–1986) 52 VACD 82% <8 cm Tumor size as prognostic factor
64% ≥8 cm
EW-87 (1987–1991) 26 Therapeutic window with IE Clinical responses in 96% Combination IE is very effective in ESFT
EW-92 (1992–1996) 34 VCDIE × 3 VCD/IE intensification 78% Tumor size (</≥8 cm) is not a prognostic factor with intensive treatment
CESS s
CESS-81 (1981–1985) 93 VACD Tumor size 80% <100 mL 31% >100 mL Tumor size (</<100 mL) and histologic response are prognostic factors
Viable tumor
79% <10%
31% <10%
CESS-86 (1986–1991) 301 SR: VACD 52% Intensive treatment with ifosfamide for high-risk patients
HR: VAID 51% (10 yr) Tumor volume >200 mL suggests poor prognosis
UKCCSG/MRC S tudies
ET-1 (1978–1986) 120 VACD 36% Tumor site is most important prognostic factor
Extr. 52%
Axial 38%
Pelvic 13%
ET-2 (1987–1993) 201 VAID 62% Extr. 73% Importance of administration of high-dose alkylating agents
Axial 55%
Pelvic 41%
EICESS s
EICESS-92 (1992–1999) 647 Randomized
SR: VAID vs VAID/VACD
HR: VAID vs EVAID
5-yr DFS:
SR: VAID 67%; VAID/VACD 68%
HR: VAID 47%; EVAID 52%
Volume (>200 mL) and histologic response as prognostic factor
CESS, Cooperative Ewing Sarcoma Study; DFS, disease-free survival; EICESS, European Intergroup Cooperative Ewing Sarcoma Study; EVAID, etoposide, vincristine, actinomycin D, ifosfamide, doxorubicin; Extr, extremity; HD, high dose; HR, high risk; IE, ifosfamide and etoposide; IESS, Intergroup Ewing Sarcoma Study; MD , moderate dose; NA, not available; POG-CCG, Pediatric Oncology Group–Children's Cancer Group; SR, standard risk; UKCCSG/MRC, United Kingdom Children's Cancer Study Group and Medical Research Council; VAC, vincristine, actinomycin D, cyclophosphamide; VACD, VAC plus doxorubicin; VAID, vincristine, actinomycin D, ifosfamide, doxorubicin; VCD, vincristine, cyclophosphamide, doxorubicin; WLI, whole-lung irradiation.

Incorporation of granulocyte colony-stimulating factor (G-CSF) into treatment regimens for many types of cancer has allowed modest dose intensification of multiple-agent chemotherapy by either increasing the total dose per cycle or shortening the time between treatments. For ESFTs, this strategy is based on use of high cumulative doses of alkylating agents and topoisomerase-II inhibitors. When these general treatment guidelines are followed, the disease-free survival rate for patients with localized, low-risk disease approaches 70% to 75%, and the OS rate for this group of patients may be greater than 80%.

The importance of dose intensification in the management of ESFT was evaluated in the second POG–CCG Ewing sarcoma trial (POG-9354/CCG-7942). In that trial, patients were randomly assigned to receive alternating courses of vincristine, doxorubicin, and cyclophosphamide with IE over either 48 or 30 weeks. The early results of that randomized trial demonstrated no difference in outcome between the standard and the dose-intensified treatment groups. An alternative to increasing dose intensity is decreasing the intervals between cycles while maintaining the same dose per cycle with the use of granulocyte colony-stimulating factor. In the United States, this is the approach taken by the Children's Oncology Group (COG) in the recently completed AEWS-0031 study, in which patients were randomly assigned to receive alternating cycles of VCD (vincristine, cyclophosphamide, doxorubicin) and IE every 3 weeks or 2 weeks, which results in 33% dose intensification. The administration of chemotherapy every 2 weeks resulted in a significantly better event free survival (73% versus 65%). This dose-compressed approach is currently considered the standard of care in the United States. Although it cannot be delivered without the use of G-CSF, it is otherwise feasible and does not result in more toxicity than the 3-week regimen.

Despite improvement, the prognosis for patients with metastatic disease continues to be very poor, and only 20% to 25% survive. Several institutions have used treatment intensification, by which very high doses of different agents are administered in a short time. In the case of ESFTs, this is a very attractive alternative because ESFTs are highly sensitive to alkylating agents, which have a steep dose–response curve. Although most patients with ESFTs treated in this manner may exhibit good clinical and histologic responses, the final results are not better than those obtained with conventional therapy.

Use of protocols that include intensification of alkylators and topoisomerase II inhibitors has resulted in a significant increase in the incidence of treatment-related leukemia—specifically, acute myelogenous leukemia (AML)—and myelodysplastic syndrome (t-AML/MDS) ( Box 92.2 ). This therapeutic strategy appears to be strongly leukemogenic, and patients are at increased risk for both alkylator-related and etoposide-related t-AML/MDS. This increased risk of t-AML/MDS appears to be related to both the increase in the total cumulative doses and dose intensification. The role of hematopoietic growth factors in this complication is unknown.

Box 92.2
St. Jude Children's Research Hospital Approach to Management of Ewing Sarcoma Family Tumors

At St. Jude Children's Research Hospital, for patients with nonmetastatic disease, regardless of the size and site of the tumor, intensive multiple-agent chemotherapy, with alternating courses of vincristine, cyclophosphamide, and doxorubicin with ifosfamide and etoposide, using a dose-compressed schedule is used. Local control is undertaken after completion of approximately 4 courses of induction chemotherapy. If the disease is considered resectable, surgery is the treatment of choice for local control, and it is always performed with curative intent. Radiation therapy is used for unresectable disease and for close or microscopic surgical margins. The same approach is followed for patients with metastatic disease to the lungs, only with addition of radiation therapy to the lungs.

For patients with bone or bone marrow metastasis, the role of intensive chemotherapy and consolidation with high-dose chemotherapy and autologous hematopoietic stem cell rescue is under ongoing evaluation. In these patients, aggressive surgical resection for local control of the primary tumor and bone metastases is performed.

All patients with recurrent disease receive second-line systemic therapy. Patients with local recurrence undergo aggressive surgery when possible, and patients with metastatic disease to the lungs receive whole-lung irradiation.

New drugs and new drug combinations continue to be investigated in the care of patients with Ewing sarcoma. Although phases I and II studies of topotecan and irinotecan as single agents have shown little or no activity in patients with refractory disease, recent studies suggest that their combination with alkylating agents may be more promising. Phase II studies of the combination of topotecan (0.75 mg/m 2 per day for 5 days) and cyclophosphamide (250 mg/m 2 per day for 5 days) resulted in responses in 36% of patients with recurrent disease and in 56% of patients with untreated metastatic disease. After demonstrating the feasibility of the incorporation of this combination in the frontline treatment of ESFTs, given in alternating cycles with VCD and IE, the COG is currently exploring the impact of this addition in a randomized manner. Also recently, the combination of irinotecan on a protracted schedule (10–20 mg/m 2 per day for 5 days in 2 consecutive weeks) with temozolomide (100 mg/m 2 per day for 5 days) has shown excellent response rates.

The use of monoclonal antibodies against IGF-1R has been explored in three phase I/II studies; in this population of patients with refractory disease, responses were seen in 10% to 20% of the patients. The COG is currently exploring the impact of the use of a monoclonal antibody against IGF-1R in combination with chemotherapy in the frontline management of patients with metastatic disease in a randomized trial.

Advances in management of ESFTs have resulted in only modest improvement in outcome among patients with metastatic disease. With appropriately intensive treatment that includes bilateral lung irradiation, however, the prognosis for patients with isolated lung metastasis may be better than for patients with extrapulmonary metastasis. For patients with bone and bone marrow metastasis therefore more intense therapies may have a role. ESFTs are very sensitive to alkylators, a group of agents with a very steep dose–response curve, which provides the basis for the use of consolidation with myeloablative therapy and autologous hematopoietic stem cell transplantation (HSCT). The results of treatment with megatherapy and HSCT for patients with high-risk ESFTs must be analyzed with caution because of the lack of randomized studies and the heterogeneity of patients and treatments. Results of most retrospective European and American studies do not seem to support the use of this approach, although there could be a role for consolidation with high-dose chemotherapy and autologous HSCT in patients with relapsed disease that show response to salvage chemotherapy. The European Euro-EWING 99 trial explored the use of consolidation with high-dose busulfan-melphalan for patients with primary disseminated ESFT; the event-free and OS rates at 3 years were 27% and 34%, respectively, suggesting that patients with primary disseminated ESFT may survive using this approach.

Improvement in outcome among patients with ESFT also must be attributed to improvement in local control, which is largely related to advances in radiation therapy and better surgical approaches. Advances in systemic therapy by means of incorporation of new drugs and treatment intensification also appear to contribute to better local control. With current multimodality intensive protocols, rates of local recurrence have decreased significantly, and little difference in efficacy has been observed between surgery and radiation therapy for local control. Surgery continues to offer slightly better results, but this observation is biased by the fact that small lesions are more likely to be managed surgically. For unresectable tumors or in case of gross residual disease, recommended doses are 55 to 60 Gy. Doses of 40 to 45 Gy are used for microscopic disease. The risk of secondary sarcoma after radiation therapy is not negligible. This risk is dose dependent, but use of lower doses of radiation therapy has been associated with higher local recurrence rates.

Neuroblastoma

Epidemiology

Neuroblastoma is the most common extracranial solid tumor of childhood, accounting for 8% to 10% of all pediatric cancers. The incidence of neuroblastoma varies in different regions of the world, with a greater prevalence in developed countries. In the United States, the annual incidence is approximately 8.2 new cases per 1 million children per year with 650 newly diagnosed cases. The median age at diagnosis is approximately 18 months. It is the most common cancer of infancy. Most (85%) of the cases are diagnosed by the age of 5 years; it is extremely rare in children older than 10 years. Familial cases occur, and a small number are inherited in an autosomal dominant manner (5%). Studies have linked germline mutations in the anaplastic lymphoma kinase (ALK) gene or in the PHOX2B gene to familial neuroblastoma. Genome-wide association studies suggest a number of germline genetic variants that may predispose to sporadic neuroblastomas. Neuroblastoma also has been found in patients with neurofibromatosis, Hirschsprung disease, Beckwith-Wiedemann syndrome, fetal hydantoin syndrome, and central hypoventilation syndrome. Microscopic neuroblast nodules are found in the adrenal glands of 2% to 3% of infants who die of nonmalignant causes before 3 months of age. It is uncertain whether these nodules represent spontaneous regression of congenital neuroblastoma or maturation of the tumor into asymptomatic benign tumors. Use of prenatal ultrasonography can coincidentally detect in utero cases as early as 31 weeks' gestation.

Mass screening programs for detection of neuroblastoma in early infancy with assays for urinary catecholamine metabolites have been conducted in Japan, Germany, and the United States. In general, patients with neuroblastoma detected in these programs have early stage disease and enjoy an otherwise excellent outcome. Mass screening, however, has not resulted in a reduction in deaths from neuroblastoma. In addition, early detection has not changed outcome among patients with tumors that have biological features associated with a poor prognosis.

Ethnic disparities have been identified with African American and Native American patients frequently presenting with high-risk features and poor outcomes.

Tumor Biology

Neuroblastoma originates in neural crest cells of the sympathetic nervous system and, not unexpectedly, secretes a variety of neurogenically derived substances, including catecholamines, neuron-specific enolase (NSE), and ferritin. Urinary excretion of abnormally high levels of catecholamine metabolites occurs in 75% to 90% of patients and is related to the degree of tumor differentiation. The metabolites most often measured in evaluation of suspected neuroblastoma are urinary vanillylmandelic acid and homovanillic acid. Levels of these catecholamines are sensitive indicators of disease status. At diagnosis, other biological markers, such as NSE, serum ferritin, and ganglioside G D2 , are present in high concentration in the serum of most patients with neuroblastoma. None of these biological markers is specific, but their presence at diagnosis may be associated with the extent of disease.

Cytogenetic studies of neuroblastoma have demonstrated segmental chromosomal abnormalities in most cases. The most common cytogenetic abnormalities in neuroblastoma are double-minute chromatin bodies (DMs), homogeneously staining regions (HSRs), and nonrandom deletion or loss of heterozygosity (LOH) of the short arm of chromosome 1p36 and 11q. DMs and HSRs both represent cytogenetic manifestations of amplified sequences of the MYCN cellular oncogene. Amplified MYCN (>10 copies per cell) is a feature in approximately 40% of neuroblastomas. It is associated with the presence of advanced disease at diagnosis and poor outcome, even in patients who have early stage disease or who come to medical attention in infancy. The DNA content of the tumor (ploidy) has been shown to correlate with therapeutic outcome and survival among infants with this tumor. Infants with hyperdiploid tumors (DNA index >1.0) have a more favorable outcome than those with diploid tumors (DNA index, 1.0). Results suggest that the presence of either MYCN amplification or diploid cellular DNA tumor content identifies whether an infant has a poor prognosis with current therapeutic regimens, independent of stage. For children older than 24 months with disseminated disease, however, the DNA content of the tumor does not appear to have prognostic importance. Recent studies suggest that additional cytogenetic imprints, such as LOH at 1p36 and 11q23, confer an aggressive phenotype. Activating mutations in the tyrosine kinase domain of the anaplastic lymphoma kinase (i) oncogene have recently been noted to account for most cases of hereditary neuroblastoma and may be somatically acquired in 5% to 10% of neuroblastoma.

Pathology

Neuroblastoma is one of the small, round, blue cell tumors of childhood. It originates from neural crest cells from within the sympathetic nervous system. In 1999, the International Neuroblastoma Pathology Classification (INPC) was established to standardize the terminology and criteria for the prognostic evaluation of the morphologic features of neuroblastic tumors in an age-linked framework. The INPC categorizes neuroblastic tumors into favorable and unfavorable histology based on the degree of neuroblastic differentiation, Schwannian stroma content, mitosis-karyorrhexis index, and age at diagnosis. The INPC distinguishes four histologic categories: neuroblastoma (Schwannian stroma poor), ganglioneuroblastoma intermixed (Schwannian stroma rich), ganglioneuroma (Schwannian stroma-dominant), and ganglioneuroblastoma nodular (composite Schwannian stroma-rich/stroma dominant and stroma poor) ( Fig. 92.8 ).

Figure 92.8, (A) Neuroblastoma with nest of dense primitive cells surrounding a fibrillar center. (B) Ganglioneuroblastoma consisting of primitive cells but with increased amounts of fibrillar material. (C) Ganglioneuroma with well-differentiated ganglion cells, Schwann cells, and nerve bundles. (D) Bone marrow involvement with metastatic neuroblastoma forming a pseudorosette.

These subgroups appear to recapitulate stages in the normal differentiation of neural crest stem cells. For example, neuroblastoma is the most primitive entity and is characterized by diffuse growth of undifferentiated neuroblastic cell nests irregularly separated by thin fibrovascular septa (see Fig. 92.8A ). By contrast, benign ganglioneuroma consists of mature ganglion cells embedded in bulky stroma composed of Schwann cell sheets enveloping neuritic processes and perineural and endoneural elements (see Fig. 92.8C ). Between these two extremes is the transitional form known as ganglioneuroblastoma (see Fig. 92.8B ). These transitional forms have been subclassified into intermixed and nodular diffuse ganglioneuroblastoma. Composite (nodular) tumor is a ganglioneuroma that contains one or more discrete nodules of pure neuroblastoma. Intermixed (diffuse) tumor contains a mixture of primitive and differentiating neuroblasts with bizarre, immature, and mature ganglion cells. Patients frequently have tumors with mixtures of these cell types, and evidence supports the clinical observation that neuroblastoma can sometimes mature into benign ganglioneuroma. Spontaneous regression and therapy-induced maturation can occur.

Clinical Manifestations

The clinical manifestations of neuroblastoma are varied. Nonspecific constitutional signs and symptoms, such as fever, general malaise, and pain, are frequent initial features. The most common sites of primary tumors are the abdomen (adrenal gland or paraspinal ganglia) and the thorax (usually the posterior mediastinum). In infants, the distribution is slightly different in that a higher proportion of primary tumors occur in the thoracic cavity than is the case in older children. Common manifestations include a hard, painless mass in the neck; a localized intrathoracic mass found incidentally on a chest radiograph; and a palpable abdominal mass, according to the location of the tumor. A palpable abdominal mass can result from an enlarging primary adrenal or retroperitoneal tumor or from hepatomegaly secondary to tumor metastasis. Children may appear chronically ill and irritable and have periorbital ecchymosis, scalp nodules, and bone pain from widespread metastasis to the bone marrow or bone ( Fig. 92.9 ; see also Fig. 92.8D ). Lower-limb paresis secondary to epidural extension of a primary paraspinal tumor growing through the intervertebral foramen can cause signs of spinal cord compression. Intermittent abdominal pain, malaise, failure to gain weight, and recurrent, unexplained fever may be present for prolonged periods before neuroblastoma is diagnosed.

Figure 92.9, (A) Periorbital ecchymosis secondary to metastatic orbital neuroblastoma. (B) Radionuclide radiolabeled m -iodobenzylguanidine (MIBG) scan shows uptake in primary adrenal neuroblastoma and skeletal metastatic disease. (C) Fluorodeoxyglucose–positron emission tomography (FDG-PET) scan shows diffuse involvement with metastatic neuroblastoma.

Seventy-five percent of patients with neuroblastoma have metastatic disease at the time of diagnosis. The most common sites of metastasis are lymph nodes (local or distant), bone marrow, liver, skin, orbit, and bone (facial bones, skull, appendicular skeleton). Nearly half of patients have widespread skeletal metastasis at diagnosis. The bones of the skull and orbit frequently are affected, so proptosis, ecchymosis, and masses beneath the scalp are frequent findings (see Fig. 92.9A ). Lung metastatic lesions are extremely uncommon when the patient is first diagnosed but can be seen after autologous HSCT. Horner syndrome (ipsilateral miosis, ptosis, and anhidrosis) may be present in patients with lesions originating in the cervical or upper thoracic sympathetic ganglia. Less frequent is the syndrome of opsomyoclonus and ataxia, manifested as acute cerebellar encephalopathy, truncal ataxia, and rapid and random conjugate eye movements (so-called dancing eyes , dancing feet ). The pathophysiological mechanism of this syndrome is unknown, but metabolic and immunologic causes have been invoked. A syndrome of chronic watery diarrhea also may occur in patients with neuroblastoma. Increased serum levels of vasoactive intestinal peptide, which lead to increased intestinal motility and secretions, have been found in some cases.

Laboratory and Radiologic Evaluation

When neuroblastoma is suspected, evaluation should be performed to establish the diagnosis, determine the extent of disease, and obtain tumor material for molecular and genetic analyses. The physical examination should include special attention to look for blood pressure abnormalities, asymmetrical pupil size, facial sweating, lower limb weakness and other signs of spinal cord compression, and evidence of increased intracranial pressure. In addition, during the clinical examination, the size of palpable masses, enlarged lymph nodes, cutaneous lesions, and the liver should be carefully documented.

Laboratory studies include complete blood cell count, renal and liver function studies, coagulation screen, urinalysis, and measurement of urine catecholamines. Serum often is obtained for LDH, NSE, ferritin, and ganglioside G D2 determinations because the results may be helpful in prognosis.

Staging of neuroblastoma requires radiologic studies and marrow aspiration and biopsy to determine the extent of local disease and distant spread. The minimal evaluation for metastatic disease includes radionuclide skeletal scintigraphy and bone marrow aspiration. CT generally has replaced intravenous pyelography, arteriography, and inferior vena cavography ( Fig. 92.10 ). Ultrasonographic studies of the abdomen and pelvis may be useful in evaluation of mass lesions and the degree of compression of vital structures causing secondary complications. MRI may further improve the accuracy of determination of the anatomic extent of disease, especially in evaluation of tumors impinging on the spinal cord and assessment for the presence of hepatic metastasis. CT- or MRI-based myelography is necessary in evaluation of all patients with neurologic evidence of cord involvement. Neuroblastoma is a tumor derived from the sympathetic nervous system, thus neuroblastoma cells typically express the norepinephrine transporter which mediates active intracellular uptake of radiolabeled m -iodobenzylguanidine (MIBG) in approximately 90% of patients. Scans performed with radionuclides such MIBG may be more accurate and sensitive for detection of nonosseous as well as osseous disease. Newer modalities such as FDG-PET combined with CT scanning may provide additional value in the management and diagnosis of MIBG-negative neuroblastoma, and in visualization of hepatic metastases and soft tissue lesions not readily visualized with MIBG scan (see Fig. 92.9B and C ). When tumor material is obtained, it should be submitted for determination of DNA content (tumor cell ploidy), detection of MYCN genomic amplification, and cytogenetic analysis and should be preserved for additional genomics evaluation.

Figure 92.10, (A) Computed tomography scan shows adrenal neuroblastoma at diagnosis. (B) Serial sections through a resected adrenal neuroblastoma show large areas of diffuse hemorrhage and calcification.

The differential diagnosis suggested by the early manifestations of the tumor is broad because the initial signs and symptoms can be so vague. The presence of an abdominal mass may suggest Wilms tumor, hydronephrotic kidney, enlarged spleen or liver, lymphoma, germ cell tumor, and mesenteric cyst. Compression of vital structures in the neck and mediastinum can cause superior vena cava syndrome indistinguishable from that caused by other tumors. Persistent diarrhea suggests the presence of a malabsorptive state. Arterial hypertension may be attributed to intrinsic renal disease or pheochromocytoma. Bone pain can simulate rheumatic fever, rheumatoid arthritis, osteomyelitis, and acute leukemia. When initial signs are attributable to widespread lymphatic metastasis, a broad range of possibilities in the differential diagnosis, including primary tumor of the lymphoreticular system, storage disease, acute infection, and primary hematologic disorder, must be considered.

The diagnosis of neuroblastoma is established by pathologic evaluation of tumor tissue obtained at biopsy or by documentation of bone marrow involvement at bone marrow trephine biopsy or aspiration with the presence of characteristic clumps or syncytia of tumor cells, together with increased urine or serum levels of catecholamines or metabolites. If the histologic diagnosis is equivocal, genetic features characteristic of neuroblastoma, such as deletion of 1p or MYCN genomic amplification, support the diagnosis.

Several staging systems have been used to classify disease extent. The Evans and POG staging systems historically were the most widely used in the United States before 1999. To facilitate comparison of clinical studies, the International Neuroblastoma Staging System (INSS), which is based on clinical and surgical evaluations, was developed by consensus of major pediatric oncology groups in the United States, Europe, and Japan. The INSS differentiates among INSS stages 1, 2A, 2B, 3, 4, and 4S based on surgical excision, lymph node involvement, and metastatic sites ( Table 92.3 ). Because the INSS is a surgically based staging system, the stage for patients with locoregional disease can vary based on degree of surgical resection. It is noteworthy that patients with localized disease who will not undergo surgery cannot be adequately staged. For these reasons, the International Neuroblastoma Risk Group (INRG) Task Force developed a new pretreatment staging system in 2005 based on clinical criteria and specific image-defined risk factors. Required imaging modalities include either CT or MRI, as well as MIBG scintigraphy. The International Neuroblastoma Risk Group Staging System (INRGSS) differentiates among L1, L2, M, and MS stage. The INRGSS differs from the INSS in that the INRGSS is based on preoperative imaging characteristics and not on surgical resection. The age for the INRGSS MS stage is set at 547 days (18 months) compared with 12 months in the INSS 4S stage. Table 92.4 compares INSS and INRGSS. The INRGSS not only allows preoperative staging but also permits reassessment during treatment.

Table 92.3
Management of Metastatic Ewing Sarcoma Family of Tumors Without Hematopoietic Stem Cell Transplantation
Study Patients ( n ) Regimen Results: 5-Year DFS Comment(s)
IESS s
IESS I-II (1975–1985) 122 VACD 30%
First POG-CCG (1988–1993) 121 A a : VACD B : VACD + IE 19% Addition of IE does not improve results
E uropean S tudies
ET-1 (1978–1986) 22 VACD 9%
ET-2 42 VAID 23%
EICESS (1990–1995) 171 VAID ± etoposide 27%
Lungs: 34%
Bone or bone marrow: 28%
Combined: 14%
Pulmonary metastasis managed with RT: 40%
I ntensification P rotocols
First POG-CCG (1988–1993) 60 C : VACD + IE 26% Intensification does not improve results
High incidence of t-AML
EW-92 19 VCDIE × 3 VCD/IE 27% Intensification does not improve results
±, With or without; DFS, disease-free survival; EICESS, European Intergroup Cooperative Ewing Sarcoma Study; IE, ifosfamide, etoposide; IESS, Intergroup Ewing Sarcoma Study; POG-CCG, Pediatric Oncology Group–Children's Cancer Group; RT, radiation therapy; t-AML, therapy-related acute myelogenous leukemia; VACD, vincristine, actinomycin D, cyclophosphamide, doxorubicin; VAID, vincristine, actinomycin D, ifosfamide, doxorubicin; VCD, vincristine, cyclophosphamide, doxorubicin.

a POG-CCG study used three different regimens: A, B, and C.

Table 92.4
Comparison Between INSS and INRGSS for Neuroblastoma
INSS INRGSS
Stage 1: localized tumor with complete gross excision; ± microscopic residual disease; representative ipsilateral lymph node negative for tumor microscopically Stage L1: localized tumor not involving vital structures as defined by IDRFs and confined to one body compartment
Stage 2A: localized tumor with incomplete gross excision; representative ipsilateral lymph node negative for tumor microscopically Stage L2: locoregional tumor with presence of one or more IDRFs
Stage 2B: localized tumor with or without complete gross excision; ipsilateral lymph node positive for tumor microscopically; enlarged contralateral lymph nodes should be negative microscopically Equals stage L2
Stage 3: unresectable unilateral tumor infiltrating across the midline; ± 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 Equals stage L2
Stage 4: any primary tumor with dissemination to distant lymph nodes, bone, bone marrow, liver, skin, or other organs Stage M: distant metastatic disease (except stage MS). Distant lymph node involvement is metastatic disease. Ascites and pleural effusion, even if malignant cells are present, do not constitute metastatic disease unless they are remote from the primary tumor.
Stage 4S: localized primary tumor in infants younger than age 1 yr (localized as in stage 1, 2A, or 2B) with dissemination limited to skin, liver, or bone marrow (<10% malignant cells) Stage MS: metastatic disease in children younger than 547 days (18 mo) of age with metastases confined to skin, liver, or bone marrow (<10% malignant cells); MIBG scan must be negative in bone and bone marrow. Primary tumor can be L1 or L2 with no limitations in terms of crossing or infiltration of the midline.
IDRF, Image-defined risk factor; INRGSS, International Neuroblastoma Risk Group Staging System; INSS, International Neuroblastoma Staging System; MIBG, 131 I- m -iodobenzylguanidine.

You're Reading a Preview

Become a Clinical Tree membership for Full access and enjoy Unlimited articles

Become membership

If you are a member. Log in here