Soft-Tissue and Bone Sarcomas

Article

Sarcomas are a heterogenous group of tumors originating from mesenchymal tissues. According to American Cancer Society estimates, approxiately 8,070 new cases will be diagnosed in 1995, including 6,000 cases of soft-tissue sarcomas and 2,070 cases of bone tumors [1].

Soft-Tissue SarcomasBone SarcomasReferences

Sarcomas are a heterogenous group of tumors originating from mesenchymal tissues. According to American Cancer Society estimates, approxiately 8,070 new cases will be diagnosed in 1995, including 6,000 cases of soft-tissue sarcomas and 2,070 cases of bone tumors [1].

Soft-Tissue Sarcomas

Soft-tissue sarcomas are extremely rare tumors. They represent 0.7% of adult malignancies [1]; it is estimated that in the United States 6,000 new cases and 3,600 deaths will occur from this disease in 1995 [1]. Soft-tissue sarcomas occur more frequently in children; they represent 6.5% of all cancers in children younger than 15 years of age [1] and are the fifth leading cause of cancer death in this age group.

Soft tissues are the extraskeletal tissues of the body that support, connect, and surround other discrete anatomic structures. These tissues contribute more than 50% of the body weight and include muscles and tendons as well as fibrous, adipose, and synovial tissues.

Soft-tissue sarcomas represent a histologically heterogeneous group of malignant tumors arising in the soft tissues. The majority of soft-tissue sarcomas are of mesodermal origin, but some sarcomas are derived from the ectoderm (eg, tumors of the connective tissues of the face and tumors composed of neurons). Although sarcomas are often thought to be exclusively mesenchymal in origin, some histologic subtypes (eg, synovial sarcoma and epithelioid sarcoma) share some epithelial features, and it is even speculated that they may be derived from epithelial tissue containing the cytokeratin type of intermediate filament [2].

Epidemiology and Pathogenesis

The pathogenesis of soft-tissue sarcomas is not completely understood. Exposure to environmental toxins has been linked with the development of two specific sarcomas: mesothelioma (asbestos [3]) and hepatic angiosarcoma (thorotrast [4] and vinyl chloride [5]). In 1979, a Swedish report linked exposure to phenoxyacetic acids (herbicides) and chlorophenols (wood preservatives)[6] to an increased risk of developing soft-tissue sarcomas, but this was not confirmed in later studies [7–9]. Ionizing radiation has been implicated as a cause of sarcomas arising in soft tissue and bone [10]. The latent period averages approximately 10 years but ranges between 2 and 30 years, and the prognosis is usually poor. The association of lymphedema and lymphangiosarcoma is well recognized, as in Stewart-Treves syndrome [11], and carries a very poor prognosis.

Despite anecdotal reports about clusters of sarcomas in some families, there is no clear genetic predisposition except in the Li-Fraumeni syndrome [12]. On the other hand, soft-tissue tumors are thought to occur more frequently in patients with a variety of genetically transmitted diseases, such as the basal cell nevus syndrome, tuberous sclerosis, Werner syndrome, intestinal polyposis, and Gardner's syndrome [13-18]. Sarcomas rarely develop from preexisting benign soft-tissue tumors [19]. The exceptions are neurofibromas in type-I Recklinghausen's disease, which have an increased risk for degeneration into malignant schwannomas. Patients with this disease have a 15% risk of developing neurofibrosarcoma and should be carefully monitored.

Recent advances in molecular biology indicate that genetic mutations in mesenchymal stem cells within the soft tissues may be responsible for the development of sarcomas. Alterations in the retinoblastoma gene and p53 gene have been found in a variety of soft-tissue sarcomas [20,21]. Germline mutations of these genes have been identified in familial retinoblastoma cases and in the Li-Fraumeni syndrome [22,23], with soft-tissue sarcomas as manifestations of these disorders. In sarcomas, mutations of p53 have been associated with specific tumor subtypes, high histologic grade, and a poor prognosis [24].

Specific cytogenetic alterations have been associated with some sarcomas and appear to be pathognomonic. For example, t(11;22) is present in 90% of patients with extraskeletal Ewing's sarcomas [25], and 50% of alveolar rhabdomyosarcomas show a t(2;13) translocation [26,27]. Myxoid liposarcomas have been found to have a t(12;16) translocation, clear-cell sarcoma a t(12;22) translocation, extraskeletal myxoid chondrosarcoma a t(9;22) translocation, and synovial sarcoma a t(x;18) translocation [28]. Neuroblastoma has been associated with structural abnormalities of chromosome 1p in 70% to 80% of cases, and these abnormalities appear to confer a poor prognosis [29]. The increasing availability of molecular biology techniques and the identification of specific DNA and RNA gene sequences as expressions of the gene product Myo D1 and oncogenes will help in the diagnosis of sarcomas.

Clinical Presentation

Soft-tissue sarcomas can occur in any anatomic region of the body because of the ubiquitous nature of connective tissue, but most sarcomas (60%) develop in the extremities. Three times as many sarcomas develop in the legs as in the arms. Other sites include the trunk (31%) and head and neck region (9%). The most common manifestations of sarcomas of the extremity are swelling and pain. Pain is usually mild and occurs later in the course of the disease. Thus, a patient might delay seeking medical attention, and a definitive diagnosis also might be delayed [31]. In children, the majority of soft-tissue sarcomas are rhabdomyosarcomas, arising in 20% of the cases in the extremities, in 37% in the head and neck region, and in 25% in genitourinary sites [32]. Patients with pelvic sarcomas might present with swelling of the leg that simulates primary iliofemoral thrombosis or with pain in the distribution of the femoral or sciatic nerve. Hypoglycemia is rare and is usually associated with large retroperitoneal sarcomas.

Evaluation

Imaging Techniques: Radiologic evaluation should include a chest x-ray and a computed tomography (CT) scan of the lungs, the most common sites of metastasis [33], and a CT scan or, preferably, a magnetic resonance imaging (MRI) scan of the primary tumor-bearing area. MRI examination of the affected area using T1-weighted images, proton-density-weighted images, and T2-weighted images can maximize the contrast among soft-tissue neoplasm, muscle, fat, and vessels [34] and almost eliminates the need for an arteriogram.

If the lesion abuts bone, a bone scan should be obtained to help determine whether there is periosteal invasion or reaction [35]. A positive bone scan does not document bone involvement by the tumor, but it may represent periosteal reaction. The bone scan can serve as a guide to wide resection near the bone or to removal of the periosteum and/or part of the bone in patients treated with surgery alone [36].

Biopsy: The biopsy of soft-tissue sarcomas is an important aspect of disease management. Needle biopsy is the preferred method because it is less invasive, less expensive, and easy to perform; however, a needle biopsy requires expertise in cytopathology for interpretation. Core biopsies can provide enough tissue for morphologic details, electron microscopy, DNA flow cytometry, cytogenetics, immunohistochemistry, and molecular studies, if necessary [37], without compromising the definitive surgery.

If an open biopsy must be performed, the biopsy site should be removed at the time of definitive resection. Therefore, it is important for the biopsy incision not to compromise subsequent surgical excision. An excisional biopsy may be used for small or superficial lesions smaller than 2 cm in diameter. The tissues surrounding the tumor form a pseudocapsule that always contains invasive prongs of malignant tissue. Therefore, shelling out soft-tissue sarcomas is never curative. Local recurrence following such procedures occurs in approximately 80% of cases.

Pathology

There are approximately 70 different histologic types of soft-tissue sarcomas. Most sarcomas are classified according to the normal cell type they mimic, based on the system proposed by Enzinger and Weiss [38]. Even among experienced pathologists, significant disagreement often arises as to the cell of origin of an individual tumor [39]. The relative frequency of the various types of sarcomas differs according to a patient's age. In children, for example, the most common sarcoma is rhabdomyosarcoma, which represents 5% to 8% of all childhood cancers. It occurs primarily in infants and children and has a predilection for the head and neck region, urinary bladder, vagina, prostate, and retroperitoneum. In older children, it can also occur in the extremities.

Rhabdomyosarcoma may present as one of four subtypes: embryonal, botryoid, alveolar, and pleomorphic. Immunohistochemistry is very helpful in the differential diagnosis; rhabdomyosarcoma stains positive with antibodies to actin, desmin, myoglobulin, S-100 protein, vimentin, and Myo-D. Electron microscopy can also help by showing the characteristic Z-band pattern of skeletal muscle differentiation [40,41]. Along with the synovial, epithelioid, and clear-cell sarcomas, rhabdomyosarcoma has a higher tendency toward lymphatic dissemination to the regional lymph nodes [42], in contrast to the generally low incidence (5%) of regional lymph-node metastases found in soft-tissue sarcomas [43].

The embryonal subtype is the most frequent and constitutes 75% of all rhabdomyosarcomas. The differential diagnosis of this particular type includes some of the “small-cell tumors,” such as lymphoma, neuroblastoma, oat-cell carcinoma of the lungs, Ewing's sarcoma, small-cell osteosarcoma, and mesenchymal chondrosarcoma.

Botryoid sarcoma is a subtype of embryonal sarcoma that has a polypoid or grapelike appearance. This tumor is commonly found in the urogenital tract of infants and children and rarely in the oral and nasal pharynges. The alveolar type-the second most common type of rhabdomyosarcoma-occurs in patients who are 10 to 25 years older and more frequently in the extremities. It is an aggressive tumor and has a poorer prognosis than the embryonal type. Pleomorphic rhabdomyosarcoma is the least common subtype. It occurs in adults, more commonly in the extremities, and should be differentiated from other pleomorphic sarcomas, pleomorphic lymphomas, pleomorphic melanomas, and carcinomas. Here again, electron microscopy and immunohistochemistry are very helpful.

In adolescents and young adults, the most common soft-tissue tumors are synovial sarcoma, epithelioid sarcoma, clear-cell sarcoma, and primitive neuroectodermal tumors. Synovial sarcoma occurs usually near the large joints of the lower extremities in patients 15 to 40 years of age and has a slightly higher incidence of lymph-node metastases [44]. Synovial sarcomas frequently calcify, which is rare for soft-tissue tumors; calcification may also occur in extraskeletal osteosarcoma, mesenchymal chondrosarcoma, and myositis ossificans.

In adults, the most common sarcoma is malignant fibrous histiocytoma (MFH), which accounts for 10% to 20% of all soft-tissue sarcomas [45]. This type of tumor is believed to be composed of neoplastic fibroblasts with acquired histiocytic features. MFH usually occurs in the thigh and retroperitoneum of adults 40 to 80 years old [46]. In general, the term refers to a high-grade sarcoma, with the exception of myxoid MFH, which is usually regarded as an intermediate-grade sarcoma.

Liposarcomas are the second most common adult sarcoma. They usually occur in the deep soft tissue of the extremities and retroperitoneum, rarely in the subcutaneous tissue, and almost never metastasize to the regional lymph nodes [47]. Liposarcomas occur slightly more frequently in men than in women (1.5:1) and can vary in behavior, ranging from low-grade, well-differentiated disease (also called atypical lipomatous tumor) to intermediate-grade myxoid liposarcoma to high-grade pleomorphic liposarcoma.

Leiomyosarcomas arise from smooth muscle and can occur anywhere in the body. They commonly arise in the retroperitoneum, where they behave very aggressively.

Neurofibrosarcomas originate from the neural sheath. They are frequently associated with Recklinghausen's disease, where patients have a 15% risk of developing neurofibrosarcomas, either de novo or from malignant transformation of preexisting benign soft-tissue tumors.

Angiosarcomas include hemangiosarcomas and lymphangiosarcomas, which arise from blood vessels and lymphatic vessels, respectively. Although they are rare, representing only 2% of all soft-tissue sarcomas, they are almost always high-grade tumors, and the 5-year survival rate of patients is only 12% despite multimodality therapy [48].

Alveolar soft-part sarcomas have no benign counterpart. These tumors evolve slowly, with most patients developing metastases that progress gradually over 5 to 15 years before death occurs. Five-year survival rates of 60% are common [49]. This subtype has a higher incidence of brain metastases.

Epithelioid sarcomas are very aggressive tumors of unknown origin. They occur almost exclusively in the extremities and have a tendency to spread to noncontiguous areas, including the skin, subcutaneous tissue, fat, bone, and draining lymph nodes, which are the most common sites of metastases for these tumors; the lungs are the second most common sites of metastases [50].

Disease Grade and Staging

Sarcomas are classified according to their grade, which represents the most important prognostic factor [51,52]. Grade 1 describes well-differentiated disease, and, at the other extreme, grade 3 refers to poorly differentiated disease. The histopathologic grade of sarcomas is based on the degree of differentiation, cellularity, number of mitoses, pleormorphism, and amount of necrosis.

In multivariate analysis, necrosis has been shown to be the best parameter for predicting prognosis [53,54]. The next important prognostic factors are the size and location of the tumor. Sarcomas located in extremities generally have a better prognosis than those not in extremities, probably because they are diagnosed earlier, are more likely to be completely resected, and have a lower risk of dissemination. The staging system of the American Joint Committee on Cancer depends largely on grade and tumor-node-metastasis (TNM) classification (Table 1)[55]. The lungs are the most frequent sites of metastasis (33%), followed by bone (23%) and the liver (15%).

T
Primary tumor
T1
Tumor smaller than 5 cm
T2
Tumor 5 cm or larger
G
Histologic grade of malignancy
G1
Low
G2
Moderate
G3
High
N
Regional lymph nodes
N0
No histologically proven regional lymph-node metastasis
N1
Histologically verified regional lymph-node metastasis
M
Distant metastases
M0
No distant metastasis
M1
Distant metastasis
IA
G1T1N0M0
IB
G1T2N0M0
IIA
G2T1N0M0
IIB
G2T2N0M0
IIIA
G3T1N0M0
IIIB
G3T2N0M0
IVA
G1-3T1-2N1M0
IVB
G1-3T1-2N0-1M1

Treatment

Local Disease: Most soft-tissue sarcomas are treated according to their grade, size, and location, except for Kaposi's sarcoma, extraskeletal Ewing's sarcoma, and rhabdomyosarcoma. In all grades of soft-tissue sarcomas, despite the presence of a pseudocapsule surrounding the tumor, extrusions of tumor extend through the capsule and form micro- and macronodules called satellites. High-grade sarcomas also have a significant incidence (almost 20%) of nodules found in normal tissue beyond the capsule-skip metastases-that are usually confined to the compartment of origin of the tumor. The goal of the treatment of local disease is local control followed by the preservation of optimal function.

The surgical margin achieved has a direct influence on the local recurrence rate [38,56-59]. Marginal en bloc excision through the pseudocapsule carries a risk of local recurrence of 70% to 90%. Wide en bloc excision through normal tissue has a local recurrence rate of 20% to 30% for low-grade lesions and as high as 50% for high-grade lesions. Radical resection involving removal of an entire compartment or amputation is associated with a risk of local recurrence of less than 50% but obvious limitations in function.

The results of combination radiotherapy (XRT) can influence decisions about proposed surgery. When properly designed XRT is coupled with surgery, narrow surgical margins with XRT have equivalent prognoses to those for wide surgical margins alone, and wide surgical margins coupled with XRT are equivalent to radical surgical margins for local control and survival [60-62]. Limb salvage should be considered when the oncologic margin is not compromised and the functional result is preferable to a prosthesis. Lymph-node dissection is not routinely recommended, even for the histologic types at high risk for lymph-node dissemination, unless clinical findings indicate that it should be performed [38,42,57].

Pre- and postoperative XRT has been used in conjunction with surgery to improve local control of the tumor. If used postoperatively, a dose of 60 to 65 Gy is required to achieve local control [60,63]. If XRT is used preoperatively, the dose required is lower (ie, 50 to 54 Gy), and the radiation field is usually smaller [64]. This is because the postoperative radiation field should include the tumor site and all tissues handled during surgery, including the stab wound and drain tube sites. Also, preoperative XRT usually causes tumor shrinkage, so the tumor is smaller at the time of surgery and has significantly fewer viable cells. Thus, the surgical margins can be safely reduced and the likelihood of reseeding tumor cells during surgery is almost eliminated.

The advantages of postoperative radiation include the feasibility of immediate surgery and the avoidance of delay in wound healing caused by prior XRT. A study conducted at The University of Texas M.D. Anderson Cancer Center showed a local failure rate of 22% for patients receiving XRT postoperatively for grades 2 and 3 tumors larger than 5 cm vs a local failure rate of 10% for patients who received preoperative treatment [65]. Comparable results are reported by other investigators [66]. Wound morbidity after surgery and XRT is adversely affected by disease located in the lower extremity, advanced patient age, and a postoperative XRT boost with an interstitial implant, as shown by multivariate analysis [67]. Accelerated fractionation was of borderline significance, whereas high pathologic grade and a resection volume of more than 200 cm³ were significant only on univariate analysis. Preoperative XRT is associated with more wound morbidity than postoperative XRT. Gentle handling of tissues during surgery with adequate hemostasis and drainage, sufficient immobilization, and omission of a postoperative boost XRT dose when possible (ie, negative histologic margins) can reduce wound morbidity [68].

Investigators at the University of California at Los Angeles [69], used intra-arterial chemotherapy with doxorubicin (Adriamycin, Rubex), 30 mg/d for 3 days as a continuous infusion, followed by preoperative XRT and wide resection, in patients with soft-tissue sarcoma. This strategy produced a local recurrence rate of less than 10%, which included an amputation rate of 5%. These results are comparable to those from more conventional surgery and radiation schemes.

Retroperitoneal sarcomas pose a complex problem, because complete resection is not often possible due to anatomic constraints [70]. In many cases, partial resection of a major organ is required, and even in completely resected sarcomas, a local recurrence rate of 50% to 70% is common [71]. XRT has been used as adjuvant therapy [72] and in unresectable disease [73]. Harrison et al reported the outcome of Yale University's experience with XRT in retroperitoneal sarcomas. All three patients who underwent a complete excision with negative margins and adjuvant XRT with more than 40 Gy survived free of disease for longer than 5 years. Among the 10 patients who underwent only a biopsy for unresectable retroperitoneal sarcomas, only 4 survived longer than 1 year, and the average radiation dose was 44 Gy. The remaining 6 patients received only an average of 27 Gy and did not survive for 1 year [72].

These results agree with the results reported by Tepper et al from Massachusetts General Hospital [73]. Of the 13 patients who had a primary tumor treated with curative intent, seven patients underwent incomplete surgical resections, three of whom had a relapse of local disease when treated with radiation doses of less than 50 Gy; the four patients with good local disease control received radiation doses varying from 4,990 to 6,070 cGy. Of the three patients with unresectable disease, one treated with 62 Gy had good local control, whereas the other two treated with less than 50 Gy had local relapses. Due to the retrospective nature of the studies and the small number of patients, adjuvant XRT for retroperitoneal sarcomas remains controversial. XRT for unresectable disease, however, does have some potential for local control and palliation.

XRT has also been used with good results in the management of desmoid tumors. Desmoid tumors lack the capacity to metastasize, but they aggressively infiltrate locally and can be fatal [74,75]. Abdominal desmoid tumors usually occur in women postpartum [76]. Patients in whom a complete resection is possible have a good prognosis. Extra-abdominal desmoid tumors usually occur in the head and neck area, shoulder girdle, and inguinal region. Wide resection is often difficult or impossible in these areas, and the recurrence rate ranges between 50% and 75% for cases involving close or positive margins [77].

XRT has been used by different investigators as an alternative treatment; it provided good local control in approximately 80% of cases when 60 Gy was given over 6 to 8 weeks [77,78-82]. A review of a 20-year experience at M.D. Anderson Cancer Center using XRT doses of 50 to 76 Gy to treat desmoid tumors revealed no evidence of a dose-control relationship, but a clear dose-complication relationship was seen. Therefore, the current recommended XRT dose at M.D. Anderson is 50 to 55 Gy at 1.8 Gy per fraction [83]. At the time of a patient's initial surgery, XRT is not routinely recommended.

A review of the M.D. Anderson experience with chemotherapy for desmoid tumors [84] identified 12 patients so treated; 10 with unresectable primary or recurrent disease and 2 treated preoperatively in an attempt to shrink the tumor and decrease surgical morbidity. Eleven patients received doxorubicin-plus-dacarbazine-based regimens; six of the nine patients whose responses could be evaluated had objective responses (two complete responses [CR] and four partial responses [PR]). With a response rate of more than 60%, chemotherapy is now recommended as primary therapy for inoperable/unresectable tumors.

Hormonal therapy with tamoxifen [85], toremifene (Estrinex) [86], and progesterone [87] has been reported to achieve long-term remission of desmoid tumors. Indomethacin and ascorbic acid have been reported to cause regression of some desmoid tumors [88]. In addition, good responses have recently been reported with a combination of methotrexate and oral etoposide (VePesid) [89].

Adjuvant Chemotherapy: Adjuvant chemotherapy is considered standard therapy for rhabdomyosarcomas [32] and extraskeletal Ewing's sarcoma [90]. These two tumors are highly responsive to systemic chemotherapy, have a high incidence of systemic micrometastases, and are associated with a 5-year disease-free survival rate of less than 20%. The role of adjuvant chemotherapy in all other soft-tissue sarcomas remains controversial.

The most active single agents for soft-tissue-sarcomas are ifosfamide (Ifex, 30%)[91], doxorubicin (26%)[92], dactinomycin (Cosmegen, 17%)[93], and dacarbazine (16%)[94]. The combination of dacarbazine and doxorubicin resulted in a response rate of 42% [95] and has been found to be superior to doxorubicin alone [96,97]. There was a dose-response relationship for doxorubicin, with regimens including doses of greater than 70 mg/m² having higher response rates than regimens using lower doses [84]; significantly less cardiotoxicity was seen when doxorubicin was administered as a prolonged infusion [99,100]. Ifosfamide (with mesna [Mesnex] for urothelial protection) resulted in a response rate of 25% to 30% in patients who did not respond to doxorubicin-based regimens [101-103]. A dose-response relationship was documented for ifosfamide in subsequent trials [103-104].

Several prospective, randomized studies were conducted to evaluate the role of adjuvant chemotherapy in localized soft-tissue sarcomas (Table 2)[105-116]. Several studies show a statistically significant improvement in disease-free survival; however, only one study (Bordeaux) showed a statistically significant difference in the rate of overall survival in favor of the adjuvant chemotherapy group. Despite criticism of the use of single-agent suboptimal-dose doxorubicin as adjuvant treatment in most of these cases and the inclusion of low-risk patients in some of the studies, results of a recent meta-analysis of 11 published prospective, randomized, adjuvant studies (that took into account only published information) revealed a disease-free survival advantage (68% vs 53%, P < .00001) and an overall survival advantage (81% vs 71%, P = .0005) for adjuvant chemotherapy for soft-tissue sarcomas [117]. A preferred approach is to provide preoperative chemotherapy to patients in a high-risk subset. This would help to identify patients who are more likely to benefit from aggressive systemic treatment while sparing the “nonresponders” from the morbidity of prolonged chemotherapy.

Study
Regimen
Follow-up
Number of patients
% Disease-free survivalª
% Disease-free survival*
% Overall survivalª
% Overall survival*
EORTC [105]
CVAD
44
468
61
61
68
71
Bordeau [106]
CVAD
40
59
37
65
43
93
MDA [107]
CVAAd
> 120
47
83
76
-
-
NCI [108]
ACM
Extremity
60
67
28
54
60
54
Trunk
36
22
47
77
61
82
Retroperitoneum
24
15
49
92
100
47
Mayo [109]
AVCAd
64
61
68
65
70
70
GOG [110]
A
60
156
45
60
47
60
Scandi [111]
A
22
139
44
40
55
52
UCLA [112]
A
28
119
52
56
70
80
ISC [113]
A
47
86
54
71
55
65
Rizzoli [114]
A
28
77
45
73
70
91
DFCI [115]
A
73
46
62
66
72
71
ECOG [116]
A
105
36
55
68
53
65

Recurrent Disease: The two most common types of disease recurrence are local recurrence and hematogenous spread that most commonly involves the lungs. Local recurrence should be treated with aggressive surgical resection or as a high-risk primary tumor, with preoperative chemotherapy followed by local therapy depending on the clinical situation. Eighty percent of local recurrences occur in the first 2 years after initial surgery, and all recurrences develop by 3 years [118]. Patients with isolated local recurrences have a 5-year survival rate of 45% to 85% when treated with aggressive local therapy [119,120].

Resection of pulmonary metastases is indicated for patients with favorable prognostic factors-a tumor-doubling time of longer than 40 days, a disease-free interval of more than 1 year, fewer than three nodules, unilateral disease, and MFH tumor histology (rather than other histologies [121-123])-yielding a 5-year survival rate of 10% to 30%, with little morbidity and very low mortality.

For patients with less favorable prognostic factors, chemotherapy is the only available treatment option, although CR rates range between 10% and 15% and only one third of patients achieve long-term disease-free survival. Doxorubicin and ifosfamide are the most active single agents, with response rates ranging from 15% to 40%. The combination of ifosfamide and doxorubicin with or without dacarbazine yielded variable results, with a CR rate of 5% to 10% and a PR rate of 25% to 48% [124-129]. Two prospective, randomized, cooperative group trials indicated better overall response rates with the ifosfamide/doxorubicin combination with or without dacarbazine but no significant improvement in CR rate or survival time [130,131].

Attempts to use high-dose chemotherapy with or without total-body irradiation (TBI) have been disappointing. Responses were of short duration, and substantial treatment-related morbidity and mortality without improvement in the survival rate were seen [132,133]. The use of growth factors, such as granulocyte colony-stimulating factor (G-CSF) and granulocyte-macrophage colony-stimulating factor (GM-CSF) allows intensification of the chemotherapeutic regimens and reduces the morbidity related to neutropenia. The multilineage growth factor PIXY 321 (a fusion protein of GM-CSF and interleukin-3) or thrombopoietin with G-CSF or GM-CSF may allow an increase in the dose intensity by counteracting dose-limiting thrombocytopenia, which currently remains a significant consideration.

Bone Sarcomas

Bone sarcomas account for 0.2% of all primary cancers in adults and approximately 5% of childhood malignancies. It is estimated that 2,070 new cases will be diagnosed in the United States in 1995 and that 1,280 deaths will be caused by bone sarcomas in the same year [1].

Bone sarcomas are classified according to the tissue of origin: I, bone-forming tumors; II, cartilage-forming tumors; III, giant-cell tumors; IV, mesenchymal tumors; and V, vascular tumors [134]. The four most common types of bone sarcomas are osteosarcoma, chondrosarcoma, MFH of bone, and Ewing's sarcoma. A team at a specialized center, including an orthopedic surgeon with experience with bone tumors, a radiologist, an experienced pathologist, and a medical oncologist, is essential for the appropriate management of these tumors. As in the case of soft-tissue sarcomas, careful preoperative evaluation is necessary for patients with bone sarcomas.

Evaluation

Imaging Techniques: Staging requires plain films and CT scans of the involved area and the lungs. CT scanning of the area, including the entire bone and adjacent joint, helps to determine intraosseous and extraosseous extension [135]. CT scans of the lungs are required to evaluate for metastatic disease because the lungs are the most common sites of metastasis via hematogenous spread [136]. Bone scans can detect bone metastases and occasional multicentric primary lesions [137,138] or skip metastases [139]. MRI films of the area are very helpful in evaluating the extent of bone-marrow involvement and in detecting skip lesions. MRI can also help to evaluate a positive bone scan when findings on the corresponding plain radiograph are normal [140,141]. Some authors believe that MRI is the best single technique to use for preoperative evaluation [142], whereas others prefer a baseline CT scan for better bony details and a preoperative MRI scan for surgical planning. As in the case of soft-tissue sarcomas, arteriograms are rarely necessary for surgical planning for bone sarcomas. However, they are very helpful in assessing the vasculature of the tumor preoperatively [143] and in determining chemotherapy response that manifests as a decrease in vascularity of the tumor [144]. After careful staging, a biopsy is performed, preferably by the same physician who will be involved in the patient's definitive treatment. The same principles apply for biopsy of bone sarcomas as for biopsy of soft-tissue sarcomas. The biopsy site should be removed during definitive resection.

Biopsy: Trephine or core biopsy under fluoroscopic visualization is recommended and in most cases (as with 89% of patients with suspected osteosarcoma at M D. Anderson [145]) yields adequate material for diagnosis. Needle biopsy has the advantage of less tissue contamination than open biopsy. If the lesion extends to the extraosseous soft tissues, a biopsy of these components should be obtained. Intraosseous lesions require perforation of the cortex. In blastic lesions, tissue should be obtained from the least dense area [146].

Staging of Bone Tumors

A surgical staging system proposed by the Musculoskeletal Tumor Society in 1980 is used for bone sarcomas (Table 3)[147]. The system is based on the fact that bone sarcomas, regardless of their histologic type, behave similarly according to grade (G), location (T), and lymph-node involvement or distant metastases (M). Bone tumors metastasize almost exclusively through hematogenous spread, with pulmonary metastases usually occurring first followed by bony involvement [148]. Lymphatic involvement occurs rarely; it is found in 10% of cases at autopsy [136] and in 3% of patients with osteosarcoma undergoing amputation. Lymphatic involvement is considered a poor prognostic sign [148].

Stage
Grade
Site
IA
Low (G1)
Intracompartmental (T1)
IB
Low (G1)
Extracompartmental (T2)
IIA
High (G2)
Intracompartmental (T1)
IIB
High (G2)
Extracompartmental (T2)
III
Any G + regional or distant metastases
Any T

Osteosarcoma

Epidemiology and Clinical Presentation: Osteosarcoma is the most common primary malignant bone tumor. It affects men more than women (ratio, 1.5:1)[1]. There is a bimodal age distribution; the first peak occurs during childhood and adolescence and the second peak occurs during the sixth decade of life. It is a high-grade, malignant, spindle-cell tumor of the bone characterized by the production of osteoid by the malignant spindle-cell stroma [149].

Genetic alterations of the retinoblastoma (Rb) gene have been identified in cases of osteosarcoma [150] in patients with hereditary retinoblastoma, who carry a risk for osteosarcoma of 7% in radiation ports and in nonirradiated long bones [151], as well as in patients with sporadic osteosarcoma [152]. Loss of heterozygosity (LOH) and DNA alterations of the Rb gene in sporadic osteosarcomas indicate a poor prognosis [153]. During childhood and adolescence, 80% to 90% of osteosarcomas occur in a lower limb [154]. The disease's peak incidence during childhood/adolescence and the predominance of the distal femur and proximal tibia as first sites of presentation in that age group indicate that osteosarcoma is associated with rapid growth of weight-bearing long bones.

In patients older than 40 years, the skull, pelvis, and mandible are frequent sites of osteosarcoma presentation. People with Paget's disease have a 10-fold risk of developing bone cancer [155]. Other conditions associated with an increased risk for bone sarcoma are hereditary multiple exostoses, enchondromatosis (Ollier's disease), enchondromatosis with skin hemangiomas (Maffucci's syndrome), polyostotic fibrous dysplasia, and osteogenesis imperfecta [156]. Ionizing radiation is the only environmental agent known to cause bone tumors. People who were exposed to radium and patients receiving radiation therapy are at increased risk for osteosarcoma [157]. Such risk is expected to decrease with the use of megavoltage therapy, which is not absorbed by bone as much as orthovoltage therapy is.

Fifty percent of all osteosarcomas occur in the knee joint area; the proximal humerus is the next most common site, with 25% of the cases. The axial skeleton is rarely involved [154]. The most common presenting symptom is pain, with a firm, palpable, soft-tissue mass fixed to the underlying bone with slight tenderness. There is no erythema or effusion in the adjacent joint, and range of motion is normal. Fewer than 1% of patients will have a pathologic fracture.

There are different types of osteosarcoma. Radiologically, the classic osteosarcoma may present as a lesion that can range from nearly normal to extremely dense or even involve complete destruction of the bone. In one series, purely lytic osteosarcoma was described in nearly 14% of patients [158]. Radiographically, purely lytic disease cannot be distinguished from a telangiectatic osteosarcoma, giant-cell tumor, aneurysmal bone cyst, or MFH of bone. The periosteal reaction classically described as “sunburst” is an important diagnostic feature; this is characteristically irregular and interrupted, which distinguishes malignancy from other benign conditions. A soft-tissue mass immediately adjacent to the bone lesion is usually present. Histologically, 75% of all osteosarcomas fall in the conventional category, which includes osteoblastic, chondroblastic, and fibroblastic subtypes. The survival rate is similar for all subtypes [159].

In patients treated with surgery alone, the size of the tumor, the age of the patient, and the degree of malignancy did not correlate with the survival rate [159]. The most significant variable was anatomic site; patients with pelvic and axial lesions had lower survival rates than patients with tumors of the extremities, probably because of incomplete resection in the former group. The preoperative serum alkaline phosphatase level has been reported as a significant prognostic factor of survival time [160]. Tumor ploidy also has been shown to be a significant prognostic factor. Disease-free survival and overall survival times are significantly longer in patients with near-diploid cell lines [161]. However, a recent review of the published data on prognostic factors in the postadjuvant-chemotherapy era assessed age, anatomic location, tumor size, and tumor necrosis following neoadjuvant chemotherapy; only tumor necrosis maintained its significance as a predictor of disease-free survival in multivariate models [162].

Treatment of Localized Disease: Prior to the 1970s, treatment of localized osteosarcoma consisted of amputation one joint above the tumor-containing bone or transmedullary amputation, with an overall survival rate of 5% to 20% at 2 years [163]. Pulmonary metastases usually occurred within 9 months and were responsible for the patient's death at 18 to 24 months after diagnosis [164]. The introduction of effective chemotherapeutic agents-high-dose methotrexate with leucovorin rescue [165], doxorubicin [166], cisplatin (Platinol)[167], and the alkylating agents ifosfamide [168] and high-dose cyclophosphamide (Cytoxan, Neosar)[169]-allowed a change in the approach to treatment of osteosarcomas.

The beneficial role of adjuvant chemotherapy in patients with osteosarcoma is now proven. A large multi-institutional study showed a 2-year disease-free survival rate of 66% for patients treated with surgery and adjuvant postoperative chemotherapy vs a 2-year disease-free survival rate of less than 20% for patients treated with surgery alone [170]. Moreover, the introduction of neoadjuvant chemotherapy allowed more conservative limb-salvage surgery in 50% to 80% of patients without compromising a patient's chance of survival by delaying surgery [171].

Currently, limb-salvage surgery is preferred for a significant number of patients with osteosarcoma and other high-grade sarcomas. The risk of local recurrence (less than 5%) was shown to be the same or lower than that of patients treated with amputation in a carefully selected group of patients [172]. Contraindications for limb-salvage surgery include major neurovascular involvement, an inappropriate biopsy site, infection, a pathologic fracture with spread of tumor cells via hematoma, extensive muscle involvement, and immature skeletal age, especially for anticipated significant discrepancies in leg length. The latter is far less critical now with the development of expandable prostheses, but limb-sparing surgery is still usually not an option for children younger than 10 years old [173].

A number of trials have evaluated the role of chemotherapy administered either preoperatively or postoperatively. Most of the chemotherapeutic regimens included combinations of doxorubicin, high-dose methotrexate, cyclophosphamide, cisplatin, and ifosfamide. Relapse-free survival rates of 48% to 77% have been reported with different regimens. The effectiveness of high-dose methotrexate, however, has not been universally accepted. Response rates ranging from 0% and 80% have been reported [174-178]; a dose-response relationship has been proposed [179].

A study from the National Cancer Institute (NCI) evaluated the relationship of dose intensity to more than 90% tumor necrosis following neoadjuvant chemotherapy and suggested that large, highly concentrated doses of doxorubicin contribute to a favorable outcome in cases of Ewing's sarcoma and osteosarcoma. Cisplatin and high-dose methotrexate also revealed significant activity [180]. Doxorubicin and cisplatin-based adjuvant chemotherapy are now accepted as standard therapy for osteosarcoma in many institutions.

The role of chemotherapy prior to surgery (neoadjuvant) vs its role postoperatively (adjuvant) has not yet been established in a randomized trial. The rationale for neoadjuvant chemotherapy has been to institute early systemic therapy against micrometastases, to increase the chances for successful limb-salvage surgery, to decrease the risk of viable tumor cells being spread at the time of surgery, and, very importantly, to tailor treatment individually according to the patient's response.

Several systems have been proposed for grading the tumor response to neoadjuvant chemotherapy, all of which are based on the degree of cellularity and necrosis in the resected specimen [181,182]. Only grade 4 response, as determined by Huvos et al (absence of viable tumor within the entire specimen)[181], and more than 90% tumor necrosis, by the M.D. Anderson criteria [182], predicted continuous disease-free survival [183].

Both intra-arterial (IA) and intravenous (IV) chemotherapy have been administered preoperatively. A randomized trial using IA vs IV cisplatin combined with systemic doxorubicin and high-dose methotrexate showed a significantly higher histologic response (more than 90% tumor necrosis) in the group treated with IA cisplatin. No differences in the percentage of clinical/radiologic response or in the percentage of limb-salvage procedures performed were seen, and no differences in local or systemic side effects were observed [184]. IA cisplatin was compared with high-dose IV methotrexate at M.D. Anderson, and IA cisplatin was found to be more efficacious [185]. It should be noted that the IA technique requires excellent angiographic support facilities and experienced personnel to minimize complications.

At the present time, preoperative chemotherapy is standard at M.D. Anderson, with four cycles of systemic doxorubicin administered as a continuous infusion followed by cisplatin, preferably given IA where feasible, followed by limb-salvage surgery and the tailoring of postoperative therapy based on the tumor's response. Therapy is switched to ifosfamide and high-dose methotrexate for patients with less than 90% tumor necrosis. Patients with more than 90% tumor necrosis receive another six cycles of doxorubicin, with dacarbazine replacing cisplatin due to dose-limiting peripheral neuropathy.

Mature data from patients treated between 1980 and 1982 (cisplatin, 120 mg/m²) have shown that patients with complete responses at the time of surgery had a 5-year continuous disease-free survival rate of 86%, compared with 13% for patients with partial and/or poor responses [183,186]. Patients treated later with intensified cisplatin (160 mg/m²) showed a slight improvement in the overall survival rate, although this difference was not statistically significant. A significant improvement in the overall survival rate was observed for patients with less than 90% tumor necrosis (33%)[187]. Late relapses, although rare, can occur; 1 of 37 patients treated at M.D. Anderson between 1980 and 1982 had a recurrence of disease more than 6 years after surgery. Therefore, careful long-term follow-up and reports of mature data are very important.

Recent studies [187a,187b] with muramyltripeptide phosphatidylethanolamine (MTP-PE), a synthetic, lipophilic analog of muramyldipeptide (MDP), the smallest component of a mycobacterium capable of stimulating the immune system, have shown that liposomes containing MTP-PE localize to the pulmonary microvasculature, resulting in activation of pulmonary macrophages. Animal studies revealed the efficacy of MTP-PE in canine osteosarcoma models [188]. A phase II study in humans documented histologic changes in the characteristics of the pulmonary nodules that recurred after treatment with MTP-PE [189]; peripheral fibrosis and inflammatory-cell infiltration with neovascularization were demonstrated as well as a change from high- to low-grade lesions after therapy. These changes indicated a potential role for MTP-PE in conjunction with surgery and chemotherapy in the treatment of osteosarcomas. A phase II trial of L-MTP-PE administered to patients with osteosarcoma who had pulmonary metastases that either developed during adjuvant chemotherapy or persisted despite chemotherapy showed a significant prolongation of disease-free survival in the patients who received treatment for 24 weeks [190].

Treatment of Metastatic Disease: Metastatic osteosarcoma is a treatable and potentially curable disease when combined modality therapy is administered. The most frequent sites of metastasis are the lungs. Whether pulmonary metastasis is present at diagnosis or occurs after the primary tumor is treated, if the lungs are the only sites of metastatic disease, it should be resected aggressively. A 5-year postthoracotomy survival rate of nearly 40% has been reported by Putnam et al [191], and similar results were reported by Skinner et al for patients treated with combined-modality therapy [192]. The number of nodules on preoperative lung tomograms, the disease-free interval, the resectability of the tumor, and the number of metastases resected at thoracotomy are prognostic factors influencing survival [191]. The completeness of the surgical resection is crucial to prolonged survival [193]. Variables such as preoperative chemotherapy vs immediate surgery, serum lactic dehydrogenase (LDH) level, alkaline phosphatase level, or the site of primary tumor did not affect the survival of patients who presented with metastatic disease [194]. Patients with unresectable metastatic disease have a poor prognosis. Ifosfamide (with mesna for urothelial protection), alone or in combination with etoposide, has shown activity in up to one third of patients with recurrent osteosarcoma [194,195].

XRT was used as primary therapy for osteosarcoma in the 1950s and early 1960s, without good results [196]. High doses of radiation were required to sterilize only a small subset of tumors and were associated with significant necrosis of normal tissue. Preoperative XRT did not offer any survival advantage either compared with surgery alone [197]. XRT has been used successfully in facial lesions; when followed by wide surgical excision, XRT promotes a 5-year survival rate of 73%, compared with 35% to 45% with surgery alone [198]. For palliation of metastatic bone sarcomas and unresectable lesions in axial sites or the pelvis, XRT is useful, especially when combined with IA or IV radiosensitizers (5-bromodeoxyuridine or idoxuridine) [199,200].

Variants of Osteosarcoma

Parosteal and periosteal osteosarcomas are the most common variants of osteosarcoma. Parosteal osteosarcoma accounts for 4% of all osteosarcomas [201]. It usually occurs in older people, with a slightly higher incidence in women. The distal femur is involved in 75% of cases. Parosteal osteosarcoma arises from the cortex of bone. It presents as a mass that is occasionally associated with pain and grows slowly over months or years with late metastases. Overall survival rates range between 75% and 85% [201,202].

Radiographically, parosteal osteosarcoma is characterized by a large, dense, lobulated mass broadly attached to the underlying bone without involvement of the medullary canal. However, such involvement can be present without being radiographically apparent. Parosteal osteosarcoma is treated surgically with wide excision alone. Dedifferentiated parosteal osteosarcoma and high-grade surface osteosarcoma have a much poorer prognosis and are treated with chemotherapy and surgery.

Periosteal osteosarcoma originates in the cortex, usually of the tibial shaft. A characteristic, small, radiolucent lesion, evidence of bone spicules on a plain radiograph, and a Codman's triangle are hallmarks of this disease [203]. Pathologically, periosteal osteosarcoma is usually a high-grade chondroblastic osteosarcoma. Treatment recommendations follow the same concepts as those for high-grade classic osteosarcoma.

Chondrosarcoma of Bone

Chondrosarcoma is the second most common primary malignant spindle-cell tumor of the bone, characterized by cartilaginous neoplastic tissue without direct osteoid formation. Occasionally, bone formation occurs from differentiated cartilage. There are five types of chondrosarcoma: central (arising within the bone), peripheral (arising from the bone surface), mesenchymal, dedifferentiated, and clear cell [204]. The most common variants are central and peripheral chondrosarcomas, which may arise as primary tumors or may be secondary to underlying benign neoplasms (multiple osteochondromas or enchondromas)[205-207]. Chondrosarcomas usually present in patients older than 40 years of age. The most common sites are the pelvis (30%), femur (20%), and shoulder girdle (15%). Chondrosarcomas usually reach a significant size before symptoms, such as a palpable mass with pain or pressure and occasionally urinary symptoms (sometimes noted with pelvic tumors), are noted.

Chondrosarcomas are pathologically classified as grade I to grade III tumors. High-grade (grade III) tumors have the worst prognosis, with a risk of metastasis of 75% [208]. Surgical removal of the tumor is the treatment for chondrosarcoma. There is no effective chemotherapy for the central and peripheral types of the disease. Dedifferentiated chondrosarcoma, on the other hand, does respond to standard chemotherapy used against osteosarcoma, and such treatment should be employed.

Few reports document the efficacy of XRT in chondrosarcomas [209-211]. Five-year local control rates ranged between 45% and 82% and were directly related to the histology of the tumor. A review of a 15-year XRT experience with chondrosarcoma of bone at M.D. Anderson revealed that none of the four patients treated with a combination of neutron- and photon-beam XRT had disease recurrence locally [211]. One of the seven patients treated with conventional radiotherapy alone experienced local disease recurrence, suggesting a benefit from the mixed-beam technique [211].

Clear-cell chondrosarcoma is the rarest type. This tumor grows slowly and locally recurs with some malignant potential. It is often confused with chondroblastoma. Treatment is wide surgical excision; systemic therapy is not required, and metastases occur only after multiple local recurrences [212].

Mesenchymal chondrosarcoma is a rare, aggressive tumor that affects younger patients. It shows a predilection for flat bones and has high metastatic potential. Treatment is wide surgical excision with adjuvant chemotherapy. XRT is used when the tumor cannot be completely excised [213,214].

Malignant Fibrous Histiocytoma of Bone

MFH is a high-grade tumor in bone as well as in soft tissues. It usually occurs during adulthood and commonly involves the metaphyseal ends of long bones, especially those of the knee joint [215]. MFH presents as a lytic, metaphyseal lesion with marked cortical disruption, minimal cortical or periosteal reaction, and no evidence of bone formation. Pathologic fractures are common [216]. The patient's alkaline phosphatase level is normal. MFH of bone disseminates rapidly to lung tissue, and lymph-node metastases have been reported in up to one third of cases with lung metastases [217].

Although data are limited, it seems that MFH of bone is sensitive to chemotherapy [218-220]. In one study, patients treated with surgery and chemotherapy had a disease-free survival rate of 59%, compared with only 5% of patients treated with surgery alone [218]. In a recently updated study, 7 of 15 patients with MFH of bone treated with neoadjuvant chemotherapy at M.D. Anderson showed more than 90% tumor necrosis. The median continuous disease-free survival for patients who achieved more than 90% tumor necrosis was 43 months, compared with only 7 months for patients with less than 90% tumor necrosis (P < .05). A trend for better overall survival was documented (66 months vs 20 months, respectively), although it was not statistically significant [221]. MFH of bone should be treated as is osteosarcoma, with chemotherapy and surgery.

Ewing's Sarcoma

Ewing's sarcoma (ES) is a rare tumor that usually occurs in bone and presents most frequently during the second decade of life; it is an unusual occurrence before 5 or after 30 years of age. In patients up to 13 years old, ES occurs with equal frequency in girls and boys; after age 13, the disease is more common in males. ES is extremely rare in African and American blacks and Chinese populations but constitutes 12% of malignant primary tumors of bone in white persons. It was originally thought to arise from endothelial cells [222], but mesenchymal, myeloid, reticular, pericystic, neuroepithelial, and primitive multipotential cells have been suspected to be the cells of origin [223]. The most widely accepted belief is that ES is of neuroectodermal origin [224]. ES is an undifferentiated, small, round-cell tumor that may be confused with other small, round, blue-cell tumors of childhood, among them small-cell osteosarcoma, mesenchymal and myxoid chondrosarcoma, rhabdomyosarcoma, lymphoma, neuroblastoma, and peripheral neuroepithelioma.

A careful review of light microscopy studies by an experienced pathologist is extremely important for the diagnosis of ES. Morphologic features may be indistinguishable from those of the peripheral primitive neuroectodermal tumors (pPNETs). Immunocytochemical staining is often positive for neuron-specific enolase (NSE)[225], although others believe that NSE is specific for pPNET and might be used to distinguish between the two entities [226]. The monoclonal antibody 5C11 has been reported to react exclusively with ES and not with pPNET [227], whereas expression of the MIC2 gene has been reported in both entities [228]. Electron microscopy studies may also assist the diagnosis, revealing dense core granules, neurites, neurotubules, and neurofilaments in prominent Golgi's complexes [229]. A chromosomal translocation t(11;22)(q24;q12) is a characteristic abnormality of ES [230,231], but it has also been reported in pPNET [232].

ES is associated with skeletal abnormalities (such as enchondroma and aneurysmal bone cyst) and genitourinary anomalies (hypospadias and duplicate collecting system).

Clinical Presentation, Prognostic Factors, and Staging: ES can affect any bone, although it most commonly presents in the femur and pelvis. The axial skeleton is often involved. In the long bones, ES usually localizes in the diaphysis, with frequent extension through the bone cortex into the soft tissues. It presents as a rapidly enlarging mass causing poorly localized pain. Constitutional symptoms, such as fatigue, weight loss, and fever, may be present, especially in metastatic disease. Leukocytosis and an elevated erythrocyte sedimentation rate (ESR) may mimic osteomyelitis. Metastases are present at the time of initial diagnosis in 15% to 50% of cases. The lungs are the most common sites of metastasis at presentation or relapse, followed by bone and bone-marrow sites. Metastases to the central nervous system occur in fewer than 1% of patients [222].

Diagnostic and staging evaluations should include plain radiographs of the involved area and the lungs, CT or MRI scans of the primary tumor, a CT scan of the lungs, a bone scan, and bone-marrow biopsy and aspirate studies. There is no uniformly accepted staging system for ES, but a TNM system seems appropriate. Lymph-node (N) involvement is rare.

Central location of the tumor, systemic symptoms at diagnosis, elevated pretreatment LDH levels [233], the presence of a gross extraosseous extension of the primary tumor [234], metastatic disease at diagnosis [235], and less-than-complete response to preoperative chemotherapy [236] are poor prognostic factors.

Treatment: ES is considered a systemic disease; even when a tumor is apparently localized, approximately 90% of the cases include occult metastatic disease. Before treatment with chemotherapy was available, local control was achieved in 50% to 85% of patients by means of surgery or radiation doses of more than 40 Gy to 50 Gy. However, only 10% of these patients survived for 5 years. Death was usually caused by distant metastatic disease [237,238]. Because XRT can establish good local control of Ewing's sarcoma [222], the role of surgery has historically been limited to diagnostic biopsy and primary control of an expendable bone, such as a rib or clavicle [239]. Surgery can also be used to treat the primary tumor when XRT would jeopardize function. However, the higher rate of local control achieved with surgery or surgery plus XRT, compared with XRT alone [240], and the desire for prevention of long-term side effects of XRT (growth failure, normal tissue damage, and development of second malignancies [241,242]) may broaden the indications for surgery.

The advent of chemotherapy significantly improved the disease-free survival rate (50% to 60%) at 2 to 3 years, compared with such rates when XRT or surgery was used alone [243-249]. The most active single agents are cyclophosphamide (50%)[250], doxorubicin (40%) [251], vincristine (Oncovin, 30%)[252], dactinomycin (33%), etoposide (30%), and high-dose melphalan (Alkeran, 80%)[253]. Adjuvant chemotherapy is now accepted as standard therapy for ES. The combination of vincristine, dactinomycin, and cyclophosphamide has been evaluated alone or with doxorubicin or bilateral pulmonary XRT in a large intergroup study [254]. According to this study, the addition of doxorubicin significantly improved relapse-free survival and overall survival. A dose-intensity relationship exists between doxorubicin and ES [180]. A high-dose, intermittent method of chemotherapy delivery using the four drugs listed yielded significantly better disease-free survival (68%) and overall survival (77%) rates than a moderate-dose, continuous method (48% and 53%, respectively)[255]. The high-dose, intermittent schedule also improved the relapse-free survival and overall survival rates of patients with ES of the pelvic or sacral bones, a group with a generally poor outcome [256].

The combination of ifosfamide and etoposide has also shown activity in newly diagnosed [257] and previously treated [258] cases of ES, with response rates of 96% and 50% to 60%, respectively. Longer follow-up is needed to assess the duration of these responses and their effect on survival rates.

The prognosis for patients with recurrent or metastatic disease is poor. Combination chemotherapy using doxorubicin, vincristine, cyclophosphamide, and dactinomycin has been the basis of treatment. Methotrexate, bleomycin (Blenoxane), and fluorouracil have been incorporated into some protocols [244,259-261], with 5-year survival rates in the 30% range. XRT to the sites of metastatic disease (bone or soft tissue) in a dose of 45 to 50 Gy can be considered for palliation. Myeloablative therapy with high-dose etoposide, fractionated high-dose melphalan with or without carboplatin (Paraplatin), and total-body irradiation has yielded promising results (a projected relapse-free survival rate of 45% at 6 years, compared with a relapse-free survival rate of 2% in the historic control group). These patients have a poor prognosis (multifocal primary tumor and early or multiple relapses) [262], and the follow-up has been short.

In contrast, Horowitz et al [263] treated poor-risk patients with TBI and autologous bone marrow transplantation support after complete remission was reached with induction therapy. However, this study failed to show any benefit over nontransplantation protocols. Therefore, the role of high-dose chemotherapy with bone marrow transplantation support remains controversial.

References:

References

1. Wingo PA, Tong T, Bolden S: Cancer statistics. CA Cancer J Clin 44:8–30, 1995.

2. Leyvraz S, Costa J: Histological diagnosis and grading of soft-tissue sarcomas. Semin Surg Oncol 4:3–6, 1988.

3. McDonald AD, McDonald JC: Epidemiology of malignant mesothelioma, in Antman KH, Aisner J (eds): Asbestos-Related Malignancy, pp 31–55. Orlando, FL, Grune & Stratton, 1987.

4. DeSilva H, Abbott J, DaMotta L, et al: Malignancy and other effects following the administration of thorotrast. Lancet II:201–204, 1965.

5. Falk H: Vinyl chloride induced hepatic angiosarocma. Int Symp Princess Takamatsu Cancer Res Fund 18:39–46, 1987.

6. Hardell L, Sandstrom A: Case-control study: Soft-tissue sarcomas and exposure to phenoxyacetic acids or chlorophenols. Br J Cancer 39:711–717, 1979.

7. Wiklund K, Holm LE: Soft tissue sarcoma risk in Swedish agricultural and forestry workers. J Natl Cancer Inst 76:229–234, 1986.

8. Greenwald P, Kovaszway B, Collins DN, et al: Sarcomas of soft tissue after Vietnam service. J Natl Cancer Inst 73:1107–1109, 1984.

9. Kang H, Enzinger FM, Breslin P, et al: Soft tissue sarcomas and military service in Vietnam: A case-control study. J Natl Cancer Inst 79:693–699, 1987.

10. Robinson E, Nerget AZ, Wylie P: Review: Clinical aspects of postirradiation sarcoma. J Natl Cancer Inst 80:233–240, 1988.

11. Stewart FW, Treves NP: Lymphangiosarcoma in postmastectomy lymphedema: A report of six cases in elephantiasis chirurgica. Cancer 1:64–81, 1948.

12. Li FP, Fraumeni JF Jr: Soft tissue sarcomas, breast cancer, and other neoplasms: A familial syndrome? Ann Intern Med 71:747–752, 1969.

13. Fraumeni JF Jr, Vogel CL, Easton JM: Sarcomas and multiple polyposis in kindred: A genetic variety of hereditary polyposis? Arch Intern Med 121:57–61, 1968.

14. Schjweisguth O, Gerard-Marchant R, Lemerle J: Naevomatose baso-cellulaire association a un rhabdomyosarcome congenital. Arch Pediatr 25:1083–1093, 1968.

15. Heard G: Malignant disease in von Recklinghausen's neurofibromatosis. Proc R Soc Med 56:502–503, 1963.

16. Reed WB, Nickel WR, Campion G: Internal manifestations of tuberous sclerosis. Arch Dermatol 87:715–728, 1963.

17. Epstein CJ, Martin GM, Schultz AL, et al: Werner's syndrome: A review of its symptomatology, natural history, pathologic features, genetics, and relationship to the natural aging process. Medicine (Baltimore) 45:177–221, 1966.

18. Usui M, Ishii S, Yamawaki S, et al: The occurrence of soft tissue sarcomas in three siblings with Werner's syndrome. Cancer 54:2580–2586, 1984.

19. Huvos AG: The spontaneous transformation of benign into malignant soft tissue tumors. Am J Surg Pathol 9(suppl):7–20, 1985.

20. Seale KS, Lange TA, Monson D, et al: Soft tissue tumors of the foot and ankle. Foot Ankle Int 9:19–27, 1988.

21. Sallourn E, Slamant F, Cailland JM, et al: Diagnostic and therapeutic problems of soft tissue tumours other than rhabdomyosarcoma in infants under 1 year of age: A clinicopathological study of 34 cases treated at the Institut Gustave-Roussy. Med Pediatr Oncol 18:37–43, 1990.

22. Malkin D, Li FP, Strong LC, et al: Germ line p53 mutations in a familial syndrome of breast cancer, sarcomas, and other neoplasms. Science 250:1233–1238, 1990.

23. Mulligan LM, Matlashewski GJ, Scrable HJ, et al: Mechanisms of p53 loss in human sarcomas. Proc Natl Acad Sci U S A 87:5863–5867, 1990.

24. Kawai A, Nogushi M, Beppu Y, et al: Nuclear immunoreaction of p53 protein in soft tissue sarcomas: A possible prognostic factor. Cancer 73:2499–2505, 1994.

25. Callen D, Smith R, Bovine A: Chromosomal analysis in Ewing's sarcoma. Pathology 99:64–66, 1987.

26. Parham DM, Shapiro DN, Downing JR, et al: Solid alveolar rhabdomyosarcomas with the t(2;13): Report of two cases with diagnostic implications. Am J Surg Pathol 18:474–478, 1994.

27. De Chiara A, T'Ang A, Triche TJ: Expression of the retinoblastoma susceptibility gene in childhood rhabdomyosarcomas. J Natl Cancer Inst 85:152–157, 1993.

28. Angervall L, Kindblom LG: Principles for pathologic-anatomic diagnosis and classification of soft-tissue sarcomas. Clin Orthop 289:9–18, 1993.

29. Sainati L, Stella M, Montaldi A, et al: Value of cytogenetics in the differential diagnosis of the small round cell tumors of childhood. Med Pediatr Oncol 20:130–135, 1990.

30. Dias P, Dilling M, Houghton P: The molecular basis of skeletal muscle differentiation. Semin Diagn Pathol 11:3–14, 1994.

31. Lawrence W Jr, Donegan WL, Natarajan N, et al: Adult soft tissue sarcomas: A pattern of care survey of the American College of Surgeons. Ann Surg 205:349–359, 1987.

32. Maurer HM, Beltangady M, Gehan EA, et al: The Intergroup Rhabdomyosarcoma Study I: A final report. Cancer 61:209–220, 1988.

33. Potter DA, Glenn J, Kinsella T, et al: Patterns of recurrence in patients with high-grade soft-tissue sarcomas. J Clin Oncol 3:353–366, 1985.

34. Greenfield GB, Arrington JA: Magnetic resonance imaging of soft tissue tumors. Cancer Control 1:581–585, 1994.

35. Enneking WF: Preoperative staging of sarcomas. Cancer Treat Symp 3:67–70, 1985.

36. Yang JC, Rosenberg SA, Glatstein EJ, et al: Sarcomas of soft tissues, in DeVita Jr VT, Hellman S, Rosenberg SA (eds): Cancer: Principles and Practice of Oncology, 4th ed, pp 1436–1488. Philadelphia, JB Lippincott Co, 1993.

37. Ball AB, Fisher C, Pittam M: Diagnosis of soft tissue tumors by Tru-cut biopsy. Br J Surg 77:756–758, 1990.

38. Enzinger FM, Weiss SW: Soft Tissue Tumors. 2nd ed. St. Louis, MO, CV Mosby Co, 1988.

39. Presant CA, Russell WO, Alexander RW, et al: Soft tissue and bone sarcoma histopathology peer review: The frequency of disagreement in diagnosis and the need for second pathology opinions: The Southeastern Cancer Study Group experience. J Clin Oncol 4:1658–1661, 1986.

40. Parham DM, Weber B, Holt H, et al: Immunohistochemical study of childhood rhabdomyosarcomas and related neoplasms: Results of the Intergroup Rhabdomyosarcoma Study Project. Cancer 67:3072–3080, 1991.

41. Dodd S, Malone M, McCulloch WM: Rhabdomyosarcoma in children: A histological and immunohistochemical study of 59 cases. J Pathol 158:13–18, 1989.

42. Weingrad DN, Rosenberg SA: Early lymphatic spread of osteogenic and soft-tissue sarcomas. Surgery 84:231–240, 1978.

43. Ruka W, Enwich LJ, Driscoll DL, et al: Prognostic significance of lymph node metastasis and bone, major vessel, or nerve involvement in adults with high-grade soft tissue sarcomas. Cancer 62:999–1006, 1988.

44. Germar R, Moore G: Synovial sarcoma. Ann Surg 131:22–25, 1975.

45. Ushigome S, Shimoda T, Nikaido T, et al: Histopathologic diagnosis and histogenetic problems in malignant soft tissue tumors: Reassessment of malignant fibrous histocytoma, epithelioid sarcoma, malignant rhabdoid tumor, and neuroectodermal tumor. Acta Pathol Jpn 42:691–706, 1992.

46. Pritchard DJ, Reiman HM, Turcotte RE, et al: Malignant fibrous histiocytoma of the soft tissues of the trunk and extremities. Clin Orthop 289:58–65, 1993.

47. Springfield D: Liposarcoma. Clin Orthop 289:50–57, 1993.

48. Holden C, Spittle M, Jones E: Angiosarcoma of face and scalp: Prognosis and treatment. Cancer 59:1046–1057, 1987.

49. Auerbach H, Brooks J: Alveolar soft part sarcoma: A clinicopathologic and immunohistochemical study. Cancer 60:66–73, 1987.

50. Chase DR, Enzinger FM: Epithelioid sarcoma. Am J Surg Pathol 9:241–263, 1985.

51. Trojani M, Contesso G, Coindre JM, et al: Soft tissue sarcomas of adults: Study of pathological prognostic variables and definition of a histopathological grading system. Int J Cancer 33:37–42, 1984.

52. Costa J, Wesley RA, Glatstein EJ, et al: The grading of soft tissue sarcomas: Results of a clinicopathologic correlation in a series of 163 cases. Cancer 53:530–541, 1984.

53. Lack EE, Steinberg SM, White DE, et al: Extremity soft tissue sarcomas: Analysis of prognostic variables in 300 cases and evaluation of tumor necrosis as a factor in stratifying higher-grade sarcomas. J Surg Oncol 41:263–273, 1989.

54. Kulander BG, Polissar L, Yang CY, et al: Grading of soft tissue sarcomas: Necrosis as a determinant of survival. Mod Pathol 2:205–208, 1989.

55. Bears OH, Henson DE, Hutter RVP, et al (eds): Manual for Staging of Cancer, American Joint Committee on Cancer, 4th ed, pp 145–149. Philadelphia, JB Lippincott Co, 1992.

56. Simon MA, Enneking WF: The management of soft tissue sarcomas of the extremities. J Bone Joint Surg (Am) 58:317–327, 1976.

57. Sugarbaker PH, Malawer MM: Musculoskeletal Survey for Cancer: Principles and Techniques. New York, NY, Thieme Medical Publishers, 1992.

58. Gerner R, Moore G, Pickren J: Soft tissue sarcomas. Ann Surg 181:803–808, 1975.

59. Contin J, McNeer G, Chu F, et al: The problems of local recurrence after treatment of soft tissue sarcomas. Ann Surg 168:47–53, 1968.

60. Suit HD, Russell WO, Martin RG: Sarcoma of soft tissue: Clinical and histopathologic parameters and response to treatment. Cancer 35:1478–1483, 1975.

61. Karakousis CP, Emrich LJ, Rao U, et al: Limb salvage in soft tissue sarcomas with selective combination of modalities. Eur J Surg Oncol 17:71–80, 1991.

62. O'Connor MI, Pritchard DJ, Gunderson LL: Integration of limb-sparing surgery, brachytherapy, and external beam irradiation in the treatment of soft-tissue sarcoma. Clin Orthop 289:73, 1992.

63. Suit HD, Mankin HJ, Wood WC, et al: Treatment of the patient with stage M0 soft tisuse sarcoma. J Clin Oncol 6:854–862, 1988.

64. Nielsen OS, Cummings B, O'Sullivan B, et al: Preoperative and postoperative irradiation of soft tissue sarcomas: Effect of radiation field size. Int J Radiat Oncol Biol Phys 21:1595–1599, 1991.

65. Barkley HT Jr, Martin RG, Romsdahl MM, et al: Treatment of soft tissue sarcomas by preoperative irradiation and conservative surgical resection. Int J Radiat Oncol Biol Phys 14:693–699, 1988.

66. Mansson E, Willems J, Aparisi T, et al: Preoperative radiation therapy of high malignancy grade soft tissue sarcoma. Acta Radiol 22:461–464, 1983.

67. Bujko K, Suit HD, Springfield DS, et al: Wound healing after preoperative radiation for sarcoma of soft tissues. Surg Gynecol Obstet 176:124–134, 1993.

68. Suit HD, Spiro IJ: The role of radiation in patients with soft tissue sarcomas. Cancer Control 1:592–598, 1994.

69. Eilber FR, Eckardt JJ, Rosen O, et al: Neoadjuvant chemotherapy and radiotherapy in the multidisciplinary management of soft tissue sarcomas of the extremity. Surg Oncol Clin North Am 2:611–620, 1993.

70. Jacques D, Coit D, Brennan M: Soft tissue sarcoma of the retroperitoneum, in Shiu MH, Brennan MF: Surgical Management of Soft Tissue Sarcoma, pp 157–169. Philadelphia, Lea & Febiger, 1989.

71. Jacques D, Coit D, Hadju S, et al: Management of primary and recurrent soft tissue sarcoma of the retroperitoneum. Ann Surg 212:51–59, 1990.

72. Harrison L, Gutierrez E, Fisher J: Retroperitoneal sarcoma: The Yale experience and review of the literature. J Surg Oncol 32:159–164, 1986.

73. Tepper J, Suit HD, Wood W: Radiation therapy of retroperitoneal soft tissue sarcoma. Int J Radiat Oncol Biol Phys 10:825–830, 1984.

74. Khorsand J, Karakousis CP: Desmoid tumors and their management. Am J Surg 149:215–218, 1985.

75. Posner MC, Shiu MH, Newsome JL, et al: The desmoid tumor: Not a benign disease. Arch Surg 124:191–196, 1989.

76. Brasfield RD, Das Gupta TK: Desmoid tumors of the anterior abdominal wall. Surgery 65:241, 1969.

77. Acker JC, Bossen EH, Halperin EC: The management of desmoid tumors. Int J Radiat Oncol Biol Phys 26:851–858, 1993.

78. Miralbell R, Suit HD, Mankin HJ, et al: Fibromatoses: From postsurgical surveillance to combined surgery and radiation therapy. Int J Radiat Oncol Biol Phys 18:535–540, 1990.

79. McCullough WM, Parsons JT, van der Griend R, et al: Radiation therapy for aggressive fibromatosis: The experience at the University of Florida. J Bone Joint Surg Am 735:717–725, 1991.

80. Karakousis CP, Mayordomo J, Zografos GC, et al: Desmoid tumors of the trunk and extremity. Cancer 72:1637–1641, 1993.

81. Bataini JP, Belloir C, Mazabraud A, et al: Desmoid tumors in adults: The role of radiotherapy in their management. Am J Surg 155:754–760, 1988.

82. Leibel SA, Wara WM, Hill DR, et al: Desmoid tumors: Local control and patterns of relapse following radiation therapy. Int J Radiat Oncol Biol Phys 9:1167–1171, 1983.

83. Sherman NE, Romsdahl MM, Evans HL, et al: Desmoid tumors: A 20-year radiotherapy experience. Int J Radiat Oncol Biol Phys 19:37–40, 1990.

84. Patel SR, Evans HL, Benjamin RS: Combination chemotherapy in adult desmoid tumors. Cancer 72:3244–3247, 1993.

85. Kinsbrunner B, Ritter S, Domingo J, et al: Remission of rapidly growing desmoid tumors after tamoxifen therapy. Cancer 52:2201–2204, 1983.

86. Brooks MD, Ebbs SR, Colletta AA, et al: Desmoid tumors treated with triphenylethylenes. Eur J Cancer 28A:1014–1018, 1992.

87. Lanari A: Effect of progesterone on desmoid tumors (aggressive fibromatosis). N Engl J Med 309:1523, 1983.

88. Waddell WR, Gerner RE: Indomethacin and ascorbate inhibit desmoid tumors. J Surg Oncol 15:85–90, 1980.

89. Weiss A, Lackman R: Chemotherapy of desmoid tumors and fibromatosis (abstract #1365). Proc Annu Meet Am Soc Clin Oncol 13:27, 1994.

90. Razek A, Perez CA, Tefft M, et al: Intergroup Ewing's sarcoma study: Local control related to radiation dose, volume, and site of primary lesion in Ewing's sarcoma. Cancer 46:516–521, 1980.

91. Klein HO, Wickramanayake PD, Coerper CL, et al: High dose ifosfamide and mesna as continuous infusionover five days-A phase I/II trial. Cancer Treat Rev 10(suppl A):167–173, 1983.

92. O'Bryan RM, Luce JK, Talley RW, et al: Phase II evaluation of Adriamycin in human neoplasia. Cancer 32:1–8, 1973.

93. Golbey R, Li MC, Kaufman RF: Actinomycin in the treatment of soft part sarcomas (abstract). James Ewing Society Scientific Program, 1968.

94. Buesa JM, Mouridsen HT, Oosterom ATV, et al: High dose DTIC in advanced soft tissue sarcomas in the adult: A phase II study of the EORTC Soft Tissue and Bone Sarcoma Group. Ann Oncol 2:307–309, 1991.

95. Gottlieb JA, Baker LH, Quagliana JM, et al: Chemotherapy of sarcomas with a combination of Adriamycin and dimethyltriazenoimidazole carboxamide. Cancer 30:1632–1638, 1972.

96. Omura GA, Major FJ, Blessing JA, et al: A randomized study of Adriamycin with and without dimethyltriazenoimidazole carboxamide in advanced uterine sarcomas. Cancer 52:626–632, 1983.

97. Borden EC, Amato DA, Rosenbaum C, et al: Randomized comparison of three Adriamycin regimens for metastatic soft tissue sarcomas. J Clin Oncol 5:840–850, 1987.

98. O'Bryan RM, Baker LH, Gottlieb JA, et al: Dose response evaluation of Adriamycin in human neoplasia. Cancer 39:1940–1948, 1977.

99. Zalupski M, Metch B, Balcerzak S, et al: Phase III comparison of doxorubicin and dacarbazine given by bolus versus infusion in patients with soft-tissue sarcomas: A Southwest Oncology Group Study. J Natl Cancer Inst 83:926–932, 1991.

100. Legha SS, Benjamin RS, Mackay B, et al: Reduction of doxorubicin cardiotoxicity by prolonged continuous intravenous infusion. Ann Intern Med 96:133–139, 1982.

101. Magrath IT, Sandlund JT, Rayner A, et al: Treatment of recurrent sarcomas with ifosfamide (abstract #C-528). Proc Annu Meet Am Soc Clin Oncol 4:136, 1985.

102. Antman KH, Ryan L, Elias A, et al: Response to ifosfamide and mesna: One hundred twenty-four previously treated patients with metastatic or unresectable sarcoma. J Clin Oncol 7:126–131, 1989.

103. Patel SR, Vadhan-Raj S, Trevino C, et al: Phase II study of high dose ifosfamide and G-CSF in patients with malignant bone tumors and metastatic soft-tissue sarcomas (abstract #1448). Proc Annu Meet Am Soc Clin Oncol 11:413, 1992.

104. Benjamin RS, Legha SS, Patel SR, et al: Single-agent ifosfamide studies in sarcomas of soft tissue and bone: The M. D. Anderson experience. Cancer Chemother Pharmacol 31(suppl 2):S174–179, 1993.

105. Bramwell V, Rouesse J, Steward W, et al: European experience of adjuvant chemotherapy for soft-tissue sarcoma: Interim report of a randomized trial of CyVADIC versus control, in Ryan JR, Baker LH (eds): Recent Concepts in Sarcoma Treatment, pp 157–164. Dordrecht, The Netherlands: Kluwer Academic, 1988.

106. Ravaud A, Nguyen BB, Coindre JM, et al: Adjuvant chemotherapy with CyVADIC in high-risk soft tissue sarcoma: A randomized prospective trial, in Salmon SE (ed): Adjuvant Therapy of Cancer VI, pp 556–566. Philadelphia, WB Saunders Co, 1990.

107. Benjamin RS, Terjanian TO, Fenoglio CJ, et al: The importance of combination chemotherapy for adjuvant treatment of high-risk patients with soft-tissue sarcomas of the extremities, in Salmon SE (ed): Adjuvant Therapy of Cancer V, pp 735–744. Orlando, FL, Grune & Stratton, 1987.

108. Chang AE, Kinsella TJ, Glatstein E, et al: Adjuvant chemotherapy for patients with high-grade soft-tissue sarcomas of the extremity. J Clin Oncol 6:1491–1500, 1988.

109. Edmonson JH, Fleming TR, Ivins JC, et al: Randomized study of systemic chemotherapy following complete excision of nonosseous sarcomas. J Clin Oncol 2:1390–1396, 1984.

110. Omura GA, Blessing JA, Major F, et al: A randomized clinical trial of adjuvant Adriamycin in uterine sarcomas: A Gynecologic Oncology Group study. J Clin Oncol 3:1240–1245, 1985.

111. Alvegard TA, Sigurdsson H, Mouridsen H, et al: Adjuvant chemotherapy with doxorubicin in high grade soft tissue sarcoma: A randomized trial of the Scandinavian Sarcoma Group. J Clin Oncol 7:1504–1513, 1989.

112. Eilber FR, Giuliano AE, Huth JF, et al: Adjuvant Adriamycin in high grade extremity soft tissue sarcomas: A randomized prospective trial. Proc Annu Meet Am Soc Clin Oncol 5:488, 1986.

113. Baker LH: Adjuvant therapy for soft tissue sarcoma, in Ryan JR, Baker LH (eds): Recent Concepts in Sarcoma Treatment, pp 144–148. Dordrecht, The Netherlands: Kluwer Academic, 1988.

114. Gherlinzoni F, Bacci G, Picci P, et al: A randomized trial for the treatment of high grade soft tissue sarcomas of the extremities: Preliminary observations. J Clin Oncol 4:552–558, 1986.

115. Antman KH, Ryan L, Borden E, et al: Pooled results from three randomized adjuvant studies of doxorubicin versus observation in soft tissue sarcomas: Ten-year results and review of the literature, in Salmon SE (ed): Adjuvant Therapy of Cancer VI, pp 529–544. Philadelphia, WB Saunders Co, 1990.

116. Antman KH, Amato D, Lerner H, et al: ECOG and Dana Farber Cancer Institute/MGH Study, in Salmon SE (ed): Adjuvant Therapy of Cancer V, pp 611–620. Orlando, FL, Grune & Stratton, 1987.

117. Zalupski M, Ryan J, Hussein M, et al: Defining the role of adjuvant chemotherapy for patients with soft tissue sarcoma of the extremities, in Salmon SE (ed): Adjuvant Therapy of Cancer VII, pp 385–392. Philadelphia, JB Lippincott Co, 1993.

118. Lindberg RD, Martin RG, Romsdahl MM, et al: Conservative surgery and postoperative radiotherapy in 300 adults with soft-tissue sarcomas. Cancer 47:2391–2397, 1981.

119. Giuliano AE, Eilber FR, Morton DL: The management of locally recurrent soft tissue sarcoma. Ann Surg 196:87–91, 1982.

120. Singer S, Antman KH, Corson JM, et al: Long term salvageability for patients with locally recurrent soft tissue sarcomas. Arch Surg 127:548–554, 1992.

121. Putman JB Jr, Roth JA, Wesley MN, et al: Analysis of prognostic factors in patients undergoing resection of pulmonary metastases from soft tissue sarcomas. J Thorac Cardiovasc Surg 87:260–268, 1984.

122. Jablons D, Steinberg SM, Roth JA, et al: Metastasectomy for soft tissue sarcoma: Further evidence for efficacy and prognostic indicators. J Thorac Cardiovasc Surg 97:695–705, 1989.

123. Casson AG, Putnam JB Jr, Natarajan G, et al: Five-year survival after pulmonary metastasectomy for adult soft tissue sarcoma. Cancer 59:662–668, 1992.

124. Bramwell V, Quirt I, Warr D, et al: Combination chemotherapy with doxorubicin and ifosfamide in advanced adult soft tissue sarcoma. J Natl Cancer Inst 81:1496–1499, 1989.

125. Shutte J, Mouridsen HT, Stewart W, et al: Ifosfamide plus doxorubicin in previously untreated patients with advanced soft tissue sarcoma. Eur J Cancer 26:558–561, 1990.

126. Loehrer PJ, Sledge GW, Nicaise C, et al: Ifosfamide plus doxorubicin in metastatic adult sarcomas: A multi-institutional phase II trial. J Clin Oncol 7:1655–1659, 1989.

127. Elias A, Ryan LM, Aisner J, et al: Mesna, doxorubicin, ifosfamide, dacarbazine (MAID) regimen for adults with advanced sarcoma. Semin Oncol 17(suppl 4):44–49, 1990.

128. Steward WP, Verzweij J, Somers R, et al: Doxorubicin plus ifosfamide with rh-GM-CSF in the treatment of advanced adult soft-tissue sarcomas: Preliminary results of a phase II study from the EORTC Soft Tissue and Bone Sarcoma Group. J Cancer Res Clin Oncol 117S:193–197, 1991.

129. Casali P, Pastorino V, Zuchinelli P, et al: Epirubicin plus ifosfamide and dacarbazine (EID) in advanced soft tissue sarcomas. Ann Oncol 3(suppl 2):S125–126, 1992.

130. Antman KH, Crowley J, Balcerzak SP, et al: An intergroup phase III randomized study of doxorubicin and dacarbazine with or without ifosfamide and mesna in advanced soft tissue and bone sarcomas. J Clin Oncol 11:1276–1285, 1993.

131. Edmonson JH, Ryan LM, Blevin RA, et al: Randomized comparison of doxorubicin alone versus ifosfamide plus doxorubicin or mitomycin, doxorubicin, and cisplatin against soft tissue sarcomas. J Clin Oncol 11:1269–1275, 1993.

132. Cheson D, Lacerna L, Leyland-Jones B, et al: Autologous bone marrow transplantation: Current status in future directions. Ann Intern Med 100:51–65, 1989.

133. Antman KH, Eder JP, Frei E: High dose chemotherapy with bone marrow support for solid tumors. Important Adv Oncol 221–235, 1987.

134. Ackerman LV, del Regato JA: Cancer Diagnosis, Treatment, and Prognosis, p 1028. St. Louis, MO, CV Mosby Co, 1954.

135. Rosenthal DI: Computed tomography in bone and soft tissue neoplasms: Application and pathologic correlation. Crit Rev Diagn Imaging 18:243–278, 1982.

136. Jeffree GM, Price CHG, Sissins HA: The metastatic spread of osteosarcoma. Br J Cancer 32:87–107, 1975.

137. Collins JD, Bassett L, Main GD, et al: Percutaneous biopsy following positive bone scans. Radiology 132:439–442, 1979.

138. McKillop JH, Etcubanas EL, Goris ML: The indications for and limitations of bone scintigraphy in osteogenic sarcoma: A review of 55 patients. Cancer 48:113–1138, 1981.

139. Malawer MM, Dunham WF: Skip metastases in osteosarcoma: Recent experience. J Surg Oncol 22:236–245, 1983.

140. Zimmer WD, Berquist TH, McLeod RA, et al: Bone tumors: Magnetic resonance imaging versus computed tomography. Radiology 155:709–718, 1985.

141. Sundaram M, McGuire MH, Herbold DR: Magnetic resonance imaging of osteosarcoma. Skeletal Radiol 16:23–29, 1987.

142. Bloem JL, Taminiau AHM, Eulderink FM, et al: Radiologic staging of primary bone sarcoma: MR imaging, scintigraphy, angiography, and CT correlated with pathologic examination. Radiology 169:805–810, 1988.

143. Hudson TM, Hass G, Enneking WF, et al: Angiography in the management of musculoskeletal tumors. Surg Gynecol Obstet 141:11–21, 1975.

144. Carrasco CH, Charnsangavel C, Raymond AK, et al: Osteosarcoma: Angiographic assessment of response to preoperative chemotherapy. Radiology 170:839–842, 1989.

145. Ayala AG, Raymond AK, Ro JY, et al: Needle biopsy of primary bone lesions: M.D. Anderson experience. Pathol Annu 24:219–251, 1989.

146. Ayala AG, Zornosa J: Primary bone tumors: Percutaneous needle biopsy: Radiologic-pathologic study of 222 biopsies. Radiology 149:675–679, 1983.

147. Enneking WF, Spanier SS, Goodman MA: A system for the surgical staging of musculoskeletal sarcoma. Clin Orthop 153:106–120, 1980.

148. McKenna RJ, Schwinn CP, Soong KY, et al: Sarcomata of the osteogenic series (osteosarcoma, fibrosarcoma, chondrosarcoma, parosteal osteosarcoma, and sarcomata) arising in abnormal bone: An analysis of 552 cases. J Bone Joint Surg Am 48:1–26, 1966.

149. Dahlin DC, Coventry MB: Osteosarcoma, a study of 600 cases. J Bone Joint Surg Am 49:101–110, 1967.

150. Togushida J, Ishizaki K, Sasaki MS, et al: Chromosomal reorganization for the expression of recessive mutation of retinoblastoma susceptibility gene in the development of osteosarcoma. Cancer Res 48:3939–3943, 1988.

151. Abramson DH, Ellsworth RM, Kitchin FD, et al: Second nonocular tumors in retinoblastoma survivors: Are they radiation-induced? Ophthalmology 91:1351–1355, 1984.

152. Togushida J, Ishizaki K, Sasaki MS, et al: Preferential mutation of paternally derived RB gene as the initial event in sporadic osteosarcoma. Nature 338:156–158, 1989.

153. Wadayama B, Togushida J, Shimizu T, et al: Mutation spectrum of the retinoblastoma gene in osteosarcomas. Cancer Res 54:3042–3048, 1994.

154. Wilner D: Osteogenic sarcoma (osteosarcoma), in Wilner D (ed): Radiology of Bone Tumors and Allied Disorders, pp 1897–2005. Philadelphia, W.B. Saunders Co, 1992.

155. Wick MB, Siegal GP, Unni KK, et al: Sarcomas of bone complicating osteitis deformans (Paget's disease), 50 years' experience. Am J Surg Pathol 5:47–59, 1981.

156. Fraumeni JF Jr, Boice JD Jr: Bone, in Schottenfeld D, Fraumeni JF Jr (eds): Cancer Epidemiology and Prevention, pp 814–826. Philadelphia, WB Saunders Co, 1982.

157. Huvos AG, Woodard HQ, Cahan WG, et al: Postradiation osteogenic sarcoma of bone and soft tissues: A clinicopathologic study of 66 patients. Cancer 55:1244–1255, 1982.

158. DeSantos LA, Edeiken B: Purely lytic osteosarcoma. Skeletal Radiol 9:1–7, 1982.

159. Lockshin MD, Higgins TT: Prognosis in osteogenic sarcoma. Clin Orthop 58:85–101, 1968.

160. Francis KC, Kohn H, Malawer MM: Osteogenic sarcoma. J Bone Joint Surg Am 55:754, 1976.

161. Look AT, Douglass EC, Meyer WH: Clinical importance of near-diploid tumor stem lines in patients with osteosarcoma of the extremity. N Engl J Med 318:1567–1572, 1988.

162. Davis AM, Bell RS, Goodwin PJ: Prognostic factors in osteosarcoma: A critical review. J Clin Oncol 12:423–431, 1994.

163. Sweetnam R: Surgical management of primary osteosarcoma. Clin Orthop 111:57–64, 1975.

164. Friedman MA, Carter SK: The therapy of osteogenic sarcoma: Current status and thoughts for the future. J Surg Oncol 45:482, 1972.

165. Jaffe N: Recent advances in the chemotherapy of metastatic osteogenic sarcoma. Cancer 30:1627–1631, 1972.

166. Cortes EP, Holland JF, Wang JJ, et al: Amputation and Adriamycin in primary osteosarcoma. N Engl J Med 291:998, 1974.

167. Ochs JJ, Freeman AI, Douglas HO Jr, et al: Cis-diamminedichloroplatinum (II) in advanced osteogenic sarcoma. Cancer Treat Rep 62:239–245, 1978.

168. Antman KH, Montella D, Rosenbaum C, et al: Phase II trial of ifosfamide with MESNA in previouly treated metastatic sarcoma. Cancer Treat Rep 69:499–504, 1985.

169. Salh RA, Graham-Pole J, Cassano W, et al: Response of osteogenic sarcoma to the combination of etoposide and cyclophosphamide as neoadjuvant chemotherapy. Cancer 65:861–865,1990.

170. Link MP, Goorin AM, Miser A, et al: The effect of adjuvant chemotherapy in patients with osteosarcoma of the extremity. N Engl J Med 314:1600–1608, 1986.

171. Eilber FR, Eckardt JJ, Morton L: Advantages in the treatment of sarcomas of the extremity: Current status of limb salvage. Cancer 54:2695–2701, 1984.

172. Murray JA: Limb salvage surgery: An overview. Cancer Bulletin 42:332–337, 1990.

173. Rosen G, Caparos B, Huvos AG, et al: Preoperative chemotherapy for osteosarcoma. Cancer 49:1221–1230, 1982.

174. Jaffe N, Farber S, Traggis D, et al: Favorable response of metastatic osteogenic sarcoma to pulsed high dose methotrexate with citrovorum rescue and radiation therapy. Cancer 31:1367–1373, 1973.

175. Pratt C, Howarth C, Ransom J, et al: High dose methotrexate used alone and in combination for measurable primary and metastatic osteosarcoma. Cancer Treat Rep 64:11–20, 1980.

176. Jaffe N, Frei E, Traggis D, et al: Weekly high dose methotrexate-citrovorum factor in osteogenic sarcoma: Presurgical treatment of primary tumor and overt pulmonary metastases. Cancer 39:45–50, 1977.

177. Edmonson J, Creagan E, Gilchrist G: Phase II study of high dose methotrexate in patients with unresectable metastatic osteosarcoma. Cancer Treat Rep 65:538–539, 1981.

178. Grem J, King S, Wittes R, et al: The role of methotrexate in osteosarcoma. J Natl Cancer Inst 80:626–655, 1988.

179. Rosen G, Nirenberg A: Chemotherapy for osteogenic sarcoma. An investigative method, not a recipe. Cancer Treat Rep 66:1687–1697, 1982.

180. Smith MA, Ungerleider RS, Horowitz ME, et al: Influence of doxorubicin dose intensity on response and outcome for patients with osteogenic sarcoma and Ewing's sarcoma. J Natl Cancer Inst 83:1460–1470, 1991.

181. Huvos AG, Rosen G, Marcove RC: Primary osteogenic sarcoma: Pathologic aspects in 20 patients after treatment with chemotherapy, en bloc resection, and prosthetic bone replacement. Arch Pathol Lab Med 101:14–18, 1977.

182. Raymond AK, Chawla SP, Carrasco CH, et al: Osteosarcoma chemotherapy effect: A prognostic factor. Semin Diagn Pathol 4:212–236, 1987.

183. Benjamin RS, Chawla SP, Murray JA, et al: Preoperative chemotherapy for osteosarcoma: A treatment approach facilitating limb salvage with major prognostic implications, in Jones SE, Salmon SE (eds): Adjuvant Therapy of Cancer IV, pp 601-610. Orlando, FL, Grune Stratton, 1984.

184. Bacci G, Picci P, Arella M, et al: Effect of intra-arterial versus intravenous cisplatin in addition to systemic Adriamycin and high dose methotrexate on histologic tumor response of osteosarcoma of the extremities. J Chemother 4:189–195, 1992.

185. Jaffe N, Raymond AK, Ayala AG, et al: Analysis of the efficacy of intra-arterial cis-diamminedichloroplatinum II and high dose methotrexate with citrovorum factor rescue in the treatment of primary osteosarcoma. Cancer Treat Rep 2:157–163, 1989.

186. Benjamin RS, Chawla SP, Carrasco CH, et al: Primary chemotherapy for osteosarcoma-implications for limb salvage and ultimate prognosis, in Jacquillat C, Weil M, Khayat D (eds): Neo-Adjuvant Chemotherapy-First International Congress, pp 557–565. London, John Libby Eurotext, 1986.

187. Benjamin RS, Patel SR, Armen T, et al: Primary chemotherapy of osteosarcoma of the extremities-Long-term follow-up (abstract #1639). Proc Annu Meet Am Soc Clin Oncol 12:470, 1993.

187a. Sone S, Fidler IJ: In vitro activation of tumoricidal properties in rat alveolar macrophages by synthetic muramyldipeptide encapsulated in liposomes. Cell Immunol 57:42–50, 1981.

187b. Key ME, Talmadge JE, Fogler WE, et al: Isolation of tumoricidal macrophages from lung melanoma metastases of mice treated systemically with liposomes containing a lipophilic derivative of muramyldipeptide. J Natl Cancer Inst 69:1189–1198, 1982.

188. MacEwen EG, Kuzzman ID, Rosenthal RC, et al: Therapy for osteosarcoma in dogs with intravenous injection of liposome-encapsulated muramyltripeptide. J Natl Cancer Inst 81:935–938, 1989.

189. Kleinerman ES, Raymond AK, Bucana C, et al: Unique histological changes in lung metastases of osteosarcoma patients following therapy with liposomal muramyltripeptide (CGP 19835A lipid). Cancer Immunol Immunother 34:211–220, 1992.

190. Kleinerman ES, Gano J, Johnson D, et al: Efficacy of liposomal muramyltripeptide (L-MTP-PE) in relapsed osteosarcoma (abstract #1445). Proc Annu Meet Am Soc Clin Oncol 12:420, 1993.

191. Putnam JB Jr, Roth JA, Wesley MN, et al: Survival following aggressive resection of pulmonary metastases from osteogenic sarcoma: Analysis of prognostic factors. Ann Thorac Surg 36:516–523, 1983.

192. Skinner KA, Eilber FR, Holmes C, et al: Surgical treatment and chemotherapy for pulmonary metastases from osteosarcoma. Arch Surg 127:1065–1071, 1992.

193. Meyers PA, Heller G, Healey JH, et al: Osteogenic sarcoma with clinically detectable metastasis at initial presentation. J Clin Oncol 11:449–453, 1993.

194. Pratt CB, Douglas EC, Etcubanas EL, et al: Ifosfamide in pediatric malignant solid tumors. Cancer Chemother Pharmacol 24(suppl 1):524–527, 1989.

195. Miser JS, Kinsella TJ, Triche TJ, et al: Ifosfamide with mesna uroprotection and etoposide: An effective regimen in the treatment of recurrent sarcomas and other tumors of children and young adults. J Clin Oncol 5(8):1191–1198, 1987.

196. Cade S: Osteogenic sarcoma: A study based on 133 patients. J R Coll Surg Edinb 1:79–111, 1955.

197. Beck JC, Wara WM, Bovill EG, et al: The role of radiation therapy in the treatment of osteosarcoma. Radiology 120:163–165, 1976.

198. Chambers RG, Mahoney WD: Osteogenic sarcoma of the mandible: Current management. Am Surg 36:463–471, 1970.

199. Martinez A, Goffinet DR, Donaldson SS, et al: Intra-arterial infusion of radiosensitizer (BUDR) combined with hypofractionated irradiation and chemotherapy for primary treatment of osteogenic sarcoma. Int J Radiat Oncol Biol Phys 2:123–128, 1985.

200. Kinsella TJ, Glatstein EJ: Clinical experience with intravenous radiosensitizers in unresectable sarcomas. Cancer 59:908–915, 1987.

201. Ahuja SC, Villacin AB, Smith J, et al: Juxtacortical (parosteal) osteogenic sarcoma. J Bone Joint Surg Am 59:632–647, 1977.

202. Unni KK, Dahlin DC, Beaubout SW, et al: Parosteal osteogenic sarcoma. Cancer 37:2466–2475, 1976.

203. Unni KK, Dahlin DC, Beaubout SW: Periosteal osteogenic sarcoma. Cancer 37:2476–2485, 1976.

204. Shives TS, Wold LE, Dahlin DC, et al: Chondrosarcoma and its variants, in Sim FH (ed): Diagnosis and Treatment of Bone Tumors: A Team Approach, pp 211–217. Thorofare, NJ, Slack, 1983.

205. Marcove RC: Chondrosarcoma: Diagnosis and treatment. Orthop Clin North Am 8:811–819, 1977.

206. Garrison RC, Unni KK, Mcleod RA, et al: Chondrosarcoma arising in osteochondroma. Cancer 49:1890–1897, 1982.

207. Marcove RC, Mike V, Hutter RVP, et al: Chondrosarcoma of the pelvis and upper end of the femur: An analysis of factors influencing survival time in one hundred and thirteen cases. J Bone Joint Surg Am 54: 561–572, 1972.

208. Sanerkin NG: The diagnosis and grading of chondrosarcoma of the bone: A combined cytologic and histologic approach. Cancer 45:582–594, 1980.

209. Austin-Seymour M, Munzenrider J, Goitein M, et al: Fractionated proton radiation therapy of chordoma and low grade chondrosarcoma of the base of the skull. J Neurosurg 70:13–17, 1989.

210. Krochak R, Harwood AR, Cummings BJ, et al: Results of radical radiation for chondrosarcoma of bone. Radiother Oncol 1: 109–115, 1983.

211. McNaney D, Lindberg RD, Ayala AG, et al: Fifteen years' radiotherapy experience with chondrosarcoma of bone. Int J Radiat Oncol Biol Phys 8:187–190, 1982.

212. Unni KK, Dahlin DC, Beaubout SW, et al: Chondrosarcoma: Clear–cell variant: A report of 16 cases. J Bone Joint Surg Am 57:676–683, 1976.

213. Harwood AR, Krajbich JI, Fornasier VL: Mesenchymal chondrosarcoma: A report of 17 cases. Clin Orthop 158:144–148, 1981.

214. Huvos AG, Rosen G, Dabska M, et al: Mesenchymal chondrosarcoma: A clinicopathologic analysis of 35 patients with emphasis on treatment. Cancer 51:1230–1237, 1983.

215. McCarthy EF, Matsuno T, Dorfman HD: Malignant fibrous histiocytoma of bone: A study of 35 cases. Hum Pathol 10:57–70, 1979.

216. Huvos AG: Primary malignant fibrous histiocytoma of bone: Clinicopathologic study of 18 patients. N Y State J Med 76:552–559, 1976.

217. Spanier SS, Enneking WF, Enriquez P: Primary malignant fibrous histiocytoma of bone. Cancer 36:2084–2098, 1975.

218. Bacci G, Springfield D, Picci P, et al: Adjuvant chemotherapy for malignant fibrous histiocytoma in the femur and tibia. J Bone Joint Surg Am 67:620–625, 1985.

219. Heeten GJ, Koops HS, Kamps WA, et al: Treatment of malignant fibrous histiocytoma of bone: A plea for primary chemotherapy. Cancer 56:37–40, 1985.

220. Earl HM, Pringle J, Kemp B, et al: Chemotherapy of malignant fibrous histiocytoma of bone (MFHB). Ann Oncol 4:409–415, 1993.

221. Patel SR, Armen T, Carrasco CH, et al: Primary chemotherapy in malignant fibrous histiocytoma of bone-Updated UT MD Anderson Cancer Center Experience, in Banzet P, Holland JF, Khayat D, et al (eds): Cancer Treatment: An Update, pp 577–580. Paris, Springer-Verlag, 1994.

222. Ewing J: Diffuse endothelioma of bone. Proc N Y Path Soc 21:17–24, 1921.

223. Yunis EJ: Ewing's sarcoma and related small round cell neoplasms in children. Am J Surg Pathol 10:54–62, 1986.

224. Cavazzana AD, Magnani JL, Ross RA, et al: Ewing's sarcoma is an undifferentiated neuroectodermal tumor, in Evans AC, D'Angio GJ, Knudson AG, et al (eds): Advances in Neuroblastoma Research II, pp 487–498. New York, Alan R Liss, 1988.

225. Pinto A, Grant LH, Hayes FA, et al: Immunohistochemical expression of neuron-specific enolase and Leu7 in Ewing's sarcoma of bone. Cancer 64:1266–1273, 1989.

226. Tsokos M, Linnoila RI, Chandra RS, et al: Neuron-specific enolase in the diagnosis of neuroblastoma and other small, round-cell tumors in children. Hum Pathol 15:575–584, 1984.

227. Hara S, Ishii E, Tanaka S, et al: A monoclonal antibody specifically reactive with Ewing's sarcoma. Br J Cancer 60:875–879, 1989.

228. Ambros IM, Ambros PF, Strehl S, et al: MIC2 is a specific marker for Ewing's sarcoma and peripheral primitive neuroectodermal tumors. Cancer 67:1886–1893, 1991.

229. Ushigome S, Shimoda T, Takaki K, et al: Immunocytochemical and ultrastructural studies of the histogenesis of Ewing's sarcoma and putatively related tumors. Cancer 64:52–62, 1989.

230. Aurias A, Rimbaut C, Buffe D, et al: Chromosomal translocation in Ewing's sarcoma. N Engl J Med 309:496–497, 1983.

231. Tur Cavel C, Aurias A, Mugneret F, et al: Chromosomes in Ewing's sarcoma: An evaluation of 85 cases and remarkable consistency of t(11;22) (q24;q12). Cancer Genet Cytogenet 321:229–238, 1988.

232. Whang-Peng J, Triche TJ, Kuntsen T, et al: Cytogenetic characterization of selected small round cell tumors of childhood. Cancer Genet Cytogenet 21:185–208, 1986.

233. Johnson RE, Pomeroy TC: Prognostic factors for survival in Ewing's sarcoma. AJR Rad Ther Nucl Med 123:598–606, 1975.

234. Mendenhall CM, Marcus RB Jr, Enneking WF, et al: The prognostic significance of soft tissue extension in Ewing's sarcoma. Cancer 51:913–917, 1983.

235. Glaubiger DL, Makuch R, Schwarz J, et al: Determination of prognostic factors and their influence on therapeutic results in patients with Ewing's sarcoma. Cancer 45:2213–2219 1980.

236. Picci P, Rougraft BT, Bacci G, et al: Prognostic significance of histopathologic response to chemotherapy in nonmetastatic Ewing's sarcoma of the extremities. J Clin Oncol 11:1763–1769, 1993.

237. Phillips RF, Higinbotham NL: The curability of Ewing's endothelioma of bone in children. J Pediatr 70:391–397, 1967.

238. Wang CC, Schultz MD: Ewing's sarcoma. N Engl J Med 248:571–576, 1953.

239. Thomas PRM, Foulkes MA, Gilula LA, et al: Primary Ewing's sarcoma of the ribs. Cancer 51:1021–1027, 1983.

240. Bacci G, Toni A, Avella M, et al: Long-term results in 144 localized Ewing's sarcoma patients treated with combined therapy. Cancer 63:1477–1486, 1989.

241. Lewis R, Marcove RC, Rosen G: Ewing's sarcoma: Functional effects of radiation therapy. J Bone Joint Surg Am 59:325–331, 1977.

242. Teft M, Lattin P, Jereb B, et al: Acute and late effects on normal tissues following combined chemotherapy and radiation therapy for childhood rhabdomyosarcoma and Ewing's sarcoma. Cancer 37:1201–1223, 1976.

243. Rosen G, Wallner N, Tau C, et al: Disease-free survival in children with Ewing's sarcoma treated with radiation therapy and adjuvant four-drug sequential chemotherapy. Cancer 33:384–393, 1974.

244. Vietti TJ, Gehan EA, Nesbit ME: Multimodal therapy in metastatic Ewing's sarcoma: An intergroup study. Natl Cancer Inst Monogr 56:279–284, 1981.

245. Rosen G, Caparros B, Nirenburg A, et al: Ewing's sarcoma: Ten-year experience with adjuvant chemotherapy. Cancer 47:2204–2213, 1981.

246. Gasparini M, Lombardi F, Gianni C, et al: Localized Ewing's sarcoma: Results of integrated therapy and analysis of failures. Eur J Cancer 17:1205–1209, 1981.

247. Hayes FA, Thompson FL, Hustu HD, et al: The response of Ewing's sarcoma to sequential cyclophosphamide and Adriamycin induction therapy. J Clin Oncol 1:45–51, 1983.

248. Kinsella TJ, Glaubiger D, Deisseroth A, et al: Intensive combined modality therapy including low-dose TBI in high-risk Ewing's sarcoma patients. Int J Radiat Oncol Biol Phys 9:1955–1960, 1983.

249. Miser J, Kinsella TJ, Triche T, et al: Preliminary results of treatment of Ewing's sarcoma of bone in children and young adults: Six months of intensive combined modality therapy without maintenance. J Clin Oncol 6:484–490, 1988.

250. Samuels ML, Howe CD: Cyclophosphamide in the management of Ewing's sarcoma. Cancer 20:961–966, 1967.

251. Oldham RK, Pomeroy TC: Treatment of Ewing's sarcoma with Adriamycin (NSC–123127). Cancer Chemother Rep 56:635–639, 1972.

252. Sutow WN: Vincristine (NSC-67574) therapy for malignant solid tumors in children (except Wilms' tumor). Cancer Chemother Rep 52:485–487, 1968.

253. Cornbleet MA, Corringham RET, Prentice HG, et al: Treatment of Ewing's sarcoma with high-dose melphalan and autologous bone marrow transplantation. Cancer Treat Rep 65:241–244, 1981.

254. Nesbit ME, Gehan EA, Burgert ED Jr, et al: Multimodal therapy for the management of primary, nonmetastatic Ewing's sarcoma of bone: A long-term follow-up of the First Intergroup Study. J Clin Oncol 8:1664–1674, 1990.

255. Burgert ED Jr, Nesbit ME, Garnsey LA, et al: Multimodal therapy for the management of nonpelvic localized Ewing's sarcoma of bone: Intergroup study: IESS-II. J Clin Oncol 8:1514–1524, 1990.

256. Evans RG, Nesbit ME, Gehan EA, et al: Multimodal therapy for the management of localized Ewing's sarcoma of pelvic and sacral bones: A report from the Second Intergoup Study. J Clin Oncol 9:1173–1180, 1991.

257. Meyer WH, Kun L, Marina N, et al: Ifosfamide plus etoposide in newly diagnosed Ewing's sarcoma of bone. J Clin Oncol 10:1737–1742, 1992.

258. Miser JS, Kinsella TJ, Triche TJ, et al: Ifosfamide with mesna uroprotection and etoposide: An effective regimen in the treatment of recurrent sarcomas and other tumors of children and young adults. J Clin Oncol 5:1191–1198, 1987.

259. Pilepich M, Vietti TJ, Nesbit ME, et al: Radiotherapy and combination chemotherapy in advanced Ewing's sarcoma: Intergroup Study. Cancer 47:1930–1936, 1981.

260. Hayes FA, Thompson EI, Parvey L, et al: Metastatic Ewing's sarcoma: Remission induction and survival. J Clin Oncol 5:1199–1204, 1987.

261. Cangir A, Vietti TJ, Gehan EA, et al: Ewing's sarcoma metastatic at diagnosis: Results and comparisons of two Intergroup Ewing's sarcoma studies. Cancer 66:887–893, 1990.

262. Burdach J, Jurgens H, Peters C, et al: Myeloablative radiochemotherapy and hematopoietic stem cell rescue in poor-prognosis Ewing's sarcoma. J Clin Oncol 11:1482–1488, 1993.

263. Horowitz ME, Kinsella TJ, Wexler LH, et al: Total-body irradiation and autologous bone marrow transplant in the treatment of high-risk Ewing's sarcoma and rhabdomyosarcoma. J Clin Oncol 11:1911–1918, 1993.

Related Videos