Start of Main Content
National Cancer Institute (NCI)

Biology of Childhood Osteogenic Sarcoma and Potential Targets for Therapeutic Development
November 29 – 30, 2001 • Bethesda, MD


Agenda  Participants  Workshop Summary

Agenda

Thursday, November 29
 
8:00 a.m. General Introduction 

Barry Anderson, Richard Gorlick 

Pathogenesis of Osteosarcoma 
8:30 a.m.  p53/SV40 
David Malkin 
8:50 a.m.  Other suppressors 
Marc Hansen 
9:10 a.m.  Cytogenetics of Osteosarcoma 
Julia Bridge 
9:30 a.m.  Molecular pathology of human osteosarcoma 
Marc Ladanyi 
9:50 a.m.  Telomerase 
Jeff Dome 
10:10 a.m.  Discussion - Mark Bernstein, David Ebb, Holcombe Grier, Neyssa Marina, Paul Meyers, Leonard Wexler 
11:10 a.m.  Cytokine/Death Receptor Pathways in Osteosarcoma 

Fas pathway 
Eugenie Kleinerman 
11:30 a.m.  Cytokine apoptosis & inhibition of angiogenesis in murine solid tumor models 
Jon Wigginton 
11:50 a.m.  Fas & TRAIL-induced cell death in ES & NB 
Maria Tsokos 
12:10 p.m.  Interferon Gamma 
Janet Houghton 
12:30 p.m.  Discussion - Mark Bernstein, David Ebb, Holcombe Grier, Neyssa Marina, Paul Meyers, Leonard Wexler 

Drug Resistance/Growth Factor Pathways in Osteosarcoma 
2:00 p.m.  p-glycoprotein 
Mark Gebhardt 
2:20 p.m.  p-glycoprotein 
Irene Andrulis 
2:40 p.m.  Discussion - Mark Bernstein, David Ebb, Holcombe Grier, Neyssa Marina, Paul Meyers, Leonard Wexler 
3:10 p.m.  Antibody based strategies for metastatic solid tumors 
Nai-Kong Cheung 
3:30 p.m.  Growth hormone 
Lee Helman 
3:50 p.m.  IGF 
Jeff Toretsky 
4:10 p.m.  Discussion - Mark Bernstein, David Ebb, Holcombe Grier, Neyssa Marina, Paul Meyers, Leonard Wexler 
4:40 p.m.  New Treatment Approaches 
6:30 p.m.  Isotopes - Samarium 
Peter Anderson 
6:50 p.m.  Inhaled GMCSF 
Carola Arndt 
7:10 p.m.  Antifolate resistance 
Richard Gorlick 
7:30 p.m.  Viral approaches 
Dale VanderPutten/Thomas Gardner 
8:20 p.m.  Slow release platinum 
Stephen Withrow 
8:40 p.m.  Antiangiogenic approaches 
Mark Kieran 
9:00 p.m.  Bisphosphonates 
John Healey 
9:20 p.m.  Discussion - Mark Bernstein, David Ebb, Holcombe Grier, Neyssa Marina, Paul Meyers, Leonard Wexler 

Friday, November 30
 
  New Target Identification in Osteosarcoma and Model Systems for Testing 
8:30 a.m.  cDNA expression arrays & oligonucleotide expression arrays 
Paul Meltzer 
8:50 a.m.  cDNA expression arrays & oligonucleotide expression arrays 
Deborah Schofield 
9:10 a.m.  Proteomics/CGH 
Ching Lau 
9:30 a.m.  Genes related to metastases 
Peter Beardsley 
10:20 a.m.  Canine models 
Chand Khanna, Stephen Withrow 
10:40 a.m.  Metastases associated genes 
Chand Khanna 
10:40 a.m.  Evaluation of camptothecins, epothilones and signaling inhibitors in osteosarcoma models 
Peter Houghton 
10:40 a.m.  Discussion - Mark Bernstein, David Ebb, Holcombe Grier, Neyssa Marina, Paul Meyers, Leonard Wexler
12:00 p.m.  Concluding Remarks

Barry Anderson/Richard Gorlick

Back to top of page
Participants

Peter Anderson, MD, PhD
Dept of Pediatrics
Mayo Clinic Rochester
200 First St SW
Rochester, MN 55905
 
Carola A S Arndt, MD
Dept of Pediatrics
Mayo Clinic Rochester
200 First St SW
Rochester, MN 55905
 
G. Peter Beardsley, MD
Dept Peds
Yale Univ
333 Cedar St
New Haven, CT 06510
 
Julia A Bridge, MD
Director, NHS Tissue Banking Core Facility
Assoc Director, Clinical Cytogenetics
Univ of Nebraska Medical Center
44th & Emile
Omaha, NE 68198
Tel: 402-559-5733
 
Jeff Dome, MD
St Jude Children's Hosp
332 North Lauderdale
Memphis, TN 38105
 
David Henry Ebb, MD
Pediatric Medicine
15 Parkman St, WAC 712
Boston, MA 02114-3117
 
Thomas A Gardner, MD
Dept of Urology, Rm RT 420
425 University Blvd
Indianapolis, IN 46202-5143
 
Holcombe E Grier, MD
Dana Farber Cancer Inst and Children's Hospital
Associate Professor, Harvard Medical School
44 Binney St
Boston, MA 02115
 
John H Healey, MD
Chief, Orthopedics
Memorial Sloan-Kettering Cancer Center
1275 York Ave
New York, NY 10021
 
Janet Hock, MD
Dept of Anatomy, Rm MS 5045L
425 University Blvd
Indianapolis, IN 46202-5143
 
Peter Houghton, PhD
St Jude Children’s Hosp
332 North Lauderdale
Memphis, TN 38105
 
Mark Kieran, MD, PhD
Dana Farber Cancer Inst
44 Binney St
Boston, MA 02115
 
Ching Ching Lau, MD, PhD
Asst Professor, Dept of Ped Hem Onc
Texas Children’s Cancer Ctr at Baylor College of Medicine
Rm TXFC 102510
6621 Fannin Stre
Houston, TX 77030-2399
 
Neyssa M Marina, MD
Prof, Pediatrics
Lucile Salter Packard Children's Hosp
SUMC, G313 HEM/ONC
Stanford, CA, 94305-5208
 
Paul Meyers, MD
Pediatrics - Box 471
Memorial Sloan-Kettering Cancer Center
1275 York Ave
New York, NY 10021
 
Jeffrey A Toretsky, MD
Ped Hem Onc
22 South Greene St, N5E16
Univ of Maryland at Baltimore
Baltimore, MD 21201
 
Dale Vanderputten, PhD
DirectGene, Inc
2661 Riva Road, Bldg 500, Suite 520
Annapolis, Md 21401
 
Jon M Wigginton, MD
Head, NCI-CCR Translational Research Initiative;
Head, Investigational Biologics Group
Pediatric Oncology Branch, NCI
Bethesda, MD 20892
 
Irene Andrulis, MD
Mt Sinai Hosp
600 University Ave Toronto, Ontario, Canada M5G 1X5
 
Sylvain Baruchel, MD
The Hospital for Sick Children
555 University Ave
Toronto, Ontario Canada M5G 1X8
 
Mark Bernstein, MD
Hopital Sainte-Justine
Hematology Oncology
3175 Cote Ste. Catherine Road
Montreal, QC H3T1C5 Canada
Canada
Nai-Kong V Cheung, MD, PhD
Pediatrics - Box 170
Memorial Sloan-Kettering Cancer Center
1275 York Ave
New York, NY 10021
 
Noel Dybdal
Genentech, Inc
Mail Stop 37, Bldg 12
460 Point San Bruno Blvd
So San Francisco, CA 94080-4990
 
Robert L Garcea, MD
Biomedical Research Bldg, Rm 811
UCHSC Box C229
4200 E Ninth Ave
Denver, CO 80262
7nbsp;
Mark Gebhardt, MD
Dana Farber Cancer Inst
44 Binney St
Boston, MA 02115
 
Mark Hansen, MD
Health Sciences Ctr, UConn
263 Farmington Avenue
Farmington, CT 06030
 
Lee J Helman, MD
Chief, Ped Oncol Branch
Div of Clin Sciences, NCI
Bldg 10, Rm 13N240
10 Center Dr, MSC 1928
Bethesda, MD 20892-1928
 
Janet A Houghton, PhD
St Jude Children’s Hosp
332 North Lauderdale
Memphis, TN 38105
 
Chand Khanna, DVM, PhD
DACVIM (Oncology)
Mail - 10/13N240 10 Center Dr
Lab - 6N10410 Center Dr
Pediatric Oncology Branch, NCI, NIH
Bethesda, MD 20892
 
Eugenie Kleinerman, MD
MD Anderson Cancer Ctr
1515 Holcombe, Box 87
Houston, TX 77030-4095
 
David Malkin, MD
The Hospital for Sick Children
555 University Ave
Toronto, Ontario Canada M5G 1X8
 
Paul Meltzer MD, PhD
Head, Section of Molecular Genetics
Cancer Genetics Branch
National Human Genome
Research Inst, NIH
49 Convent Dr MSC 4465
Bethesda, MD 20892-4465
 
Dr Deborah Schofield
Dept of Path/CHLA
4650 Sunset Blvd
Los Angeles, CA 90054-0700
 
Maria Tsokos, MD
National Institutes of Health
Bldg 10, Rm 2A-10
9000 Rockville Pike
Bethesda, Maryland 20892
 
Leonard Wexler, MD
Pediatrics - Box 210
Memorial Sloan-Kettering Cancer Center
1275 York Ave
New York, NY 10021
 
Stephen J Withrow, PhD
Chief, Clin Oncology Service
A212 Vet. Teaching Hosp
Colorado State Univ
Fort Collins CO 80523
970-491-4175

Back to top of page
Meeting Minutes

This multidisciplinary meeting was convened on November 29-30, 2001, at the Bethesda Marriott (Bethesda, MD), to address issues concerning the biology of childhood osteogenic sarcoma and the potential targets for therapeutic development. Various members of the scientific community from different clinical and laboratory-based disciplines were invited to participate.

General Introduction - Barry D. Anderson, M.D., Ph.D., and Richard Gorlick, M.D.

Dr. Richard Gorlick opened the meeting by addressing the goals and general objectives of the meeting as follows:
•   To review the current state of knowledge and share information regarding osteosarcoma biology
•   To identify, prioritize, and support the development of biology studies of potential clinical relevance in osteosarcoma
•   To discuss appropriate methods for analysis of osteosarcoma samples
•   To discuss tissue resources that may be available for the conduct of osteosarcoma biology studies (e.g., P9851 – Osteosarcoma Biology Protocol)

In the past, most osteosarcoma biology studies have been single institution-based and have studied a single gene or pathway. Dr. Gorlick described P9851 as a single biology study that serves as a companion to all osteosarcoma clinical studies and encompasses the work of numerous laboratories. The major accomplishment of this protocol has been the accrual of specimens within the entire Cooperative Group setting; however, the assays being performed are based on the interests of the individual investigators. Finite amounts of tissue exist, so the inclusion of new, additional assays may require the elimination of current assays. As the biological investigation is broadened, the resources are likely to remain finite, so less tissue may be available for all studies. Dr. Gorlick asked whether analyzing tissue was the most relevant study due to the possible artifact of sample heterogeneity and stromal contamination. He also asked if the correct sample procurement techniques were being utilized and whether there were any alternatives to the current techniques.

Dr. Gorlick concluded his introductory talk by listing the general objectives of each session as follows:
•   The potential clinical relevance of the biological work presented (potential prognostic factor, therapeutic target, complements clinical trial directions, etc.)
•   Whether the current laboratory effort dedicated to a particular biological investigation thus far is sufficient, excessive, or insufficient
•   The most appropriate method(s) that should be utilized to pursue the biological investigation
•   Identify specific plans for the conduct of research related to the biological topic

Pathogenesis of Osteosarcoma

p53/SV40 – David Malkin, M.D.

Dr. Malkin described work with p53 and SV40. Dr. Malkin specializes in the role of germ-line p53 mutations in cancer predisposition and potential association of p53 and oncogenic viruses. He described the major functions of the p53 tumor suppressor gene as:

  1. modulation of cell cycle arrest at G1/S
  2. induction of apoptosis
  3. assisting in DNA repair
  4. inhibition of blood vessel formation (antiangiogenesis)

P53 is activated through different pathways including:

  1. DNA damage
  2. aberrant growth signal from oncogenes
  3. chemotherapeutic agents, UV light, nucleoside depletion, or protein kinase inhibitors (e.g. ATR and casein kinase II)

There are different mechanisms of inactivating p53 in human cancers, but his focus has been on mutations in the p53 gene, and viral infection. Dr. Malkin gave an overview of adenoviral-mediated p53 gene therapy. The rationale for investigating p53 gene therapy in osteosarcoma includes the knowledge that metastatic osteosarcoma is refractory to conventional therapy and that this drug-resistance can be associated with p53 mutations. He has hypothesized, as tested in four osteosarcoma cell lines (SaOS-2, HOS, KHOS/NP, MNNG), that the introduction of wild-type p53 into osteosarcoma cell lines inhibits cell growth, induces apoptosis, and sensitizes cells to chemotherapeutic agents. The studies concluded that the efficiency of adenoviral infection varies among the four cell lines in a dose-dependent manner. Ad-wtp53 (wild type p53), but not Ad-mutp53 or Ad-lacZ, treatment causes a decrease in cell viability and suppresses cell growth in a dose-dependent manner, induces apoptosis, and sensitizes osteosarcoma cell lines to cisplatin and, to a lesser degree, adriamycin. Future studies will:

  1. determine changes in levels of Bax, p21, and VEGF after treatment (using ELISA and Western blot)
  2. measure cell cycle changes after treatment (using FACS)
  3. test the p53-induction system with other sarcomas
  4. consider development of an in vivo model to examine the therapeutic potential of adenoviral p53 introduction, especially in metastatic lung disease (osteosarcoma tends to metastasize to the lung, and the adenovirus appears to have an affinity for respiratory tissue)

Dr. Malkin then discussed SV40, a DNA polyomavirus that consists of three major non-structural proteins: the small t antigen (itself is not sufficient to transform cells, but enhances the ability of large T antigen to transform cells), the smaller t antigen (function unknown), and the large T antigen (assists in viral replication, interacts with p53 and Rb, and inhibits p53 function, thus promoting cell proliferation). A 1996 study concluded that 11 of 18 osteosarcoma samples showed SV40 (Carbone, et al, Oncogene 13:527, 1996); a 1998 study showed no correlation of presence of SV40 with p53 or Rb mutation status (Mendoza et al, Oncogene 17:2457-62, 1998); and a 1997 study showed 50% of osteosarcomas tested showed SV40 in each of four regions of the viral genome (Lednicky, et al, Int. J. Cancer 72:791-800, 1997). Although SV40 is not observed in all tumors, it may exist in select families. It is postulated that SV40 large T antigen may inactivate the wild-type p53 allele in some cell types of susceptible individuals, resulting in tumor development. In 2 families with Li-Fraumeni syndrome, in which family members with cancer harbored germline heterozygous p53 mutations, SV40 large T antigen was amplified and expressed in tumors in which the wild-type p53 allele was retained. These included osteosarcoma and choroid plexus carcinoma, but not in rhabdomyosarcoma (Malkin et al, Oncogene 20: 4441-4449, 2001). In conclusion, Dr. Malkin expressed the need for: prospective studies of presence of SV40 in newly diagnosed patients; serologic analysis using newly developed ELISA assays that can differentiate between closely related potential etiologic viruses; PCR analysis using multiple primer sets of both tumor material and peripheral blood (buffy coat); Southern blot of material with adequate DNA; and correlation of virus status and other gene markers (e.g. p53/Rb).

Other Suppressors - Marc Hansen, Ph.D.

Dr. Hansen examined 32 osteosarcoma specimens to discern predictors of favorable outcome. He described a highly altered genome as common in osteosarcoma, and a number of genes that show LoH (3q, 13q, 17p, 18q). RB1, p53, and 3q have been suggested as having some predictability for outcome.

He looked further at 3q and identified EDR3, which is homologous to genes in the polycomb family. EDR3 binds to condensing chromatin. The polycomb genes regulate pRB by regulating p16. In a cancer model, EDR3 shows a shift in reversible growth arrest and irreversible growth arrest, and it is hypothesized that EDR3 regulates this shift. If this function is lost, p16 cannot be regulated and a tumor develops.

Dr. Hansen’s recent work with chromosome 18q (also associated with Paget’s Disease) examined RANK (a TNF-alpha family member) in the small region, 18q21.3. Various tests have confirmed that a normal copy of RANK is overexpressed in the nucleus of osteosarcoma cells. RANK activates the NFKappa-B pathway. RANKL (RANK Ligand) was also expressed in the nucleus, but was not overexpressed. RANKL is in the same region as RANK, but is usually found in the cytoplasm of normal cells. OPG expression in osteosarcoma cell lines was also nuclearized without being overexpressed. The RANK/RANKL is normal for growth stimulation and the decoy is OPG. RANK/RANKL also co-localize in the primary tumor nucleus; no other TNF-alpha does that. TRAIL, another TNF-alpha family member, was not found in the nucleus of osteosarcoma cells.

RANK is displaced to the cytoplasm with Ohs50 hybrids (need to define Ohs50), thus there is an upregulation and a nuclearization of RANK and TRAIL in osteosarcoma. Tumorigenesis also seems to correlate with nuclearization. Dr. Hansen discussed a model that demonstrates an upregulation of RANK and TRAIL, and nuclearization of RANK, RANK1, and OPG during tumorigenesis. More research is needed to decipher these findings.

Cytogenetics of Osteosarcoma - Julia A. Bridge, M.D.

Dr. Bridge began by stating that fresh tissue is needed for proper culture. Karotyping is both diagnostically useful and provides direction for molecular studies. Although karotyping is very difficult, the advantages do outweigh the limitations. A University of Nebraska Medical Center study analyzed 128 osteosarcoma specimens, of which 70% were clonally abnormal. Most are complex karotypes and have pronounced cell-to-cell variation or heterogeneity within the same patient. There are marker chromosomes in the majority of osteosarcomas (58%), and ring chromosomes are common. There is also evidence of genomic amplification (hsr or dmin) in at least 1/3 of cases.

Dr. Bridge recently collaborated with Swedish researchers to examine chromosome 12. Amplification of MDM2 was found in all low-grade and some high-grade tumors. In contrast to the low grade osteosarcoma, however, there was also an overrepresentation of chromosome 12p sequences in high-grade tumors, which may represent a pathway to aggressive osteosarcoma.

Cytogenetic abnormalities can be numerical or structural. Common numerical abnormalities in osteosarcoma include: gain of chromosome 1, loss of chromosome 9, 10, 13, and/or 17, and partial or complete loss of long arm of chromosome 6. Structural abnormalities include: certain common chromosomal regions that show structural rearrangements on chromosomes 11 and 19.

Dr. Bridge concluded by stating that despite the karyotypic complexity observed in osteosarcoma specimens, it is evident that specific chromosomal bands and/or regions are nonrandomly involved in osteosarcoma and may provide useful clinical information. Additional molecular investigations are needed to address the involvement of these recurrently involved chromosomes and chromosomal breakpoints.

Q/A Session

Q to Dr. Bridge: Were the high-grade tumors she studied originally low-grade tumors?

A: The tumors presented as high-grade tumors.

Q to Dr. Bridge: Were the karotypes representative of osteosarcoma in general?

A: They were representative of osteosarcoma in general.

Q to Dr. Bridge: Could a karotype be derived from every osteosarcoma sample?

A: She responded that when there is a lot of necrosis, the chance of successfully obtaining a result is less likely. The elapsed time between when the specimen is surgically removed and then received in the laboratory can also influence the ability to successfully obtain results. Notably, the ability to obtain cytogenetic results does not appear to be dependent on whether or not it is a high-grade or low-grade tumor.

Molecular Pathology of Human Osteosarcoma - Marc Ladanyi, M.D.

Dr. Ladanyi presented a synthesis of available data on the molecular pathology of human osteosarcoma. A model of the pathogenesis of osteosarcoma from 1999 proposed that most osteosarcomas have some type of combined inactivation of the RB and p53 tumor suppressor pathways, but the timing or sequence of these alterations remains unclear. Amplification of the 12q13 region (containing MDM2 and CDK4) or INK4A deletion can affect both pathways, and indeed these alterations seldom coexist with RB or P53 alterations. Because these tumors almost universally show evidence of different types of genetic alterations inactivating the RB and p53 tumor suppressor pathways, it would be expected that individual alteration types would not be efficient at identifying prognostic subsets. Indeed, new data show low prognostic significance of p53 mutations in sporadic osteosarcoma. In a recent study (Wunder, et al; CTOS meeting 11/01), 22% showed p53 mutations, but there was no impact on distant recurrence. Another interesting finding was that p53 status was concordant in all paired samples of primary and distant metastases, further supporting p53 pathway alteration as an early, pre-clinical event in osteosarcoma pathogenesis.

It may be useful to place osteosarcoma in the context of other sarcomas. There are two major classes of sarcomas:

  1. sarcomas, such as osteosarcoma, with complex unbalanced karotypes and alterations of p53 pathway in most cases (karotypic complexity indicates major telomere dysfunction in pathogenesis)
  2. sarcomas with balanced karotypes (reciprocal translocations) and alterations of p53 pathway in relatively few cases, such as Ewing's Sarcoma (suggesting a lack of a major role for telomere dysfunction in their biology)

Dr. Ladanyi discussed telomere function, which is just beginning to be investigated in osteosarcoma. Telomere shortening limits the growth and reproductive capacity of human cells. Telomere erosion has a proposed dual role in tumorigenesis including: limiting tumorigenesis when telomere erosion acts as a tumor suppressor; and promoting tumorigenesis, when impaired p53 function allows telomere erosion to progress to a late crisis (a.k.a. "genetic catastrophe").

There appears to be an interrelationship between telomere function, p53, and karotype instability. A premalignant stage includes proliferation of cells and shortening of telomeres, and only cells that lack p53 function are able to survive this stage. In the absence of p53 function, chromosomal instability develops. Therefore, either p14ARF loss or p53 mutation may allow the pre-neoplastic cell to proceed to late crisis.

Complex unbalanced karotypes are of significance in osteosarcoma. Karotypic complexity reflects chromosomal fusion-bridge-breakage cycles that occur due to advanced telomere erosion. This etiologic scenario may be similar to that found in epithelial cancers. Thus any model of the molecular pathology of human osteosarcoma needs to integrate the emerging data on telomere dysfunction.

The 2001 model of osteosarcoma pathogenesis suggests there must be an inappropriate proliferative signal that affects the osteoblast, or precursor cells, followed by loss of cell cycle control leading to early crisis. If this early crisis is followed by telomere dysfunction, karotype instability, and selection of p53 inactivation, then a late crisis occurs. Finally, selection of telomerase activation and stabilization of an abnormal karotype leads to fully malignant osteosarcoma.

Future needs and directions include: moving from lists of genetic alterations to functionally related groups of genetic alterations (hyperproliferative, cell cycle control, apoptosis, DNA damage response); a better understanding of the timing and relationship of common oncogenetic events in osteosarcoma (alterations of p53 and RB as early preclinical events, p53 pathway alterations in osteosarcoma in familial retinoblastoma; RB pathway alterations in osteosarcoma in Li-Fraumeni syndrome); developing a comprehensive analysis of the p53 and RB pathways in a single large set of osteosarcoma by new high-throughput approaches; a better understanding of different “equivalent” common oncogeneic events (preferential 12q13 amplification in low-grade/surface osteosarcoma, preferential p53 missense mutation in adult osteosarcoma); a better understanding of the paradox of carcinoma-type cytogenetics in the setting of a younger age range; and defining the biologic/genetic subsets according to telomerase status and karotypic complexity (e.g. approx. ½ of osteosarcoma are telomerase-positive and the other ½ are telomerase-negative and show alternative lengthening of telomeres).

Telomerase - Jeffrey S. Dome, M.D.

Dr. Dome discussed telomeres and the functions they serve, which include:

  1. preventing chromosomes from being recognized as damaged DNA
  2. preventing end-to-end fusions and recombinations
  3. accommodating the loss of DNA that occurs with each round of replication

It is hypothesized that normal somatic cells have finite proliferative capacity, and that telomere length is one of the checkpoints that determines whether a cell stops dividing. As cells divide, the telomere length gradually decreases to a critical size, at which point senescence is triggered. Rare cells bypass this checkpoint by inactivating the Rb or p53 pathways, in which case they continue to divide until telomeres become very short and apoptosis is triggered. These two checkpoints can be overcome by activating mechanisms to lengthen telomeres; a central feature of cancer cells is that they must come up with a way to maintain telomeres. About 85% of all cancers activate an enzyme called telomerase to maintain telomeres and the other 15% use a recombination-based method called ALT (alternative telomere lengthening). Unlike most cancers, at least 50% of osteosarcomas are dependent upon the ALT mechanism to maintain telomeres.

The ALT vs. telomerase mechanisms are a means to the same end, but they are not equivalent. This is evident in the length of the telomere itself. Telomerase-dependent osteosarcoma cell lines have short telomeres and a minor range of length, whereas ALT-dependent osteosarcoma cell lines have long telomeres and great variability in telomere length. The ALT cell lines have greater genetic instability and more translocations than the telomerase-positive cell lines. These findings are based on only 3 cell lines, but the information is suggestive. It has been hypothesized that ALT-dependent osteosarcomas have different clinical behavior than telomerase-dependent osteosarcomas.

Telomerase inhibition may represent a promising adjunct to conventional therapy. To test the theory that telomerase inactivation leads to arrest of cellular division and apoptosis, mutant forms of telomerase were developed. These mutant forms of telomerase resulted in shortening of the telomeres and eventual apoptosis. The cells with regular telomerase activity continued to grow. This suggests that a telomerase inhibitor could inhibit cellular proliferation and telomerase may prove to be a therapeutic target. However, it can take numerous cell doublings for growth arrest and apoptosis to occur, so a telomerase-inhibitory approach may be most efficacious in the setting of minimal residual disease (MRD). Emerging data indicates that telomere shortening is associated with increased radiosensitivity. Therefore, telomerase inhibition may be used with other treatments to increase the effectiveness of therapy.

Limitations of telomerase inhibition in osteosarcoma include:

  1. 50% of osteosarcomas are telomerase-negative
  2. there is unclear potential for telomerase-dependent tumors to "convert" to ALT dependent tumors
  3. substantial tumor growth may occur before telomerase inhibition takes effect

Future directions include: determining the frequency of ALT versus telomerase- dependent osteosarcomas; comparing clinical features (stage, response rate, RFS) between ALT- and telomerase-dependent osteosarcomas; comparing molecular phenotype between ALT- and telomerase-dependent osteosarcomas; and assessing telomerase inhibitors in telomerase-positive osteosarcomas.

Discussion - David H. Ebb, M.D.

Q to Dr. Dome: What is the doubling time of the cell lines that he investigated?

A: HELA cells have rapid doubling time. Cells died in about a month, but it varies depending on cell type.

Q to Dr. Hansen: Was phosphorylation modified?

A: Phosphorylation does get modified. The working hypothesis is that there are a variety of targets such as SKb1, n-myc, etc. He is currently investigating osteosarcoma cell line microarrays.

Q to Dr. Dome: Regarding the ALT pathway, how much is it used in the development of stem cells?

A: We know stem cells have telomerase activity as well as malignant cells. Telomerase inhibition may not be as selective as originally believed, but stem cells aren’t always dividing, so non-dividing stem cells would not be affected by telomerase inhibition.

Q to Dr. Malkin: Is there a pharmaceutical application for adenoviral p53 introduction on the horizon ?

A: Yes, adenoviral-based p53 vectors have been developed, and some are in early clinical trials, as well as in preclinical testing.

Q to Dr. Hansen: Questions about RANK and its potential clinical significance. Is this related to tumor aggressiveness?

A: There hasn’t been a large enough survey to say anything about this clinically and about outcome. RANK is being studied in breast cancer that has metastasized to bone.

Q to Dr. Hansen: Where would RANK be placed in the proposed model?

A: Based on frequency, mechanisms of p53 and RB are first, and RANK occurs soon after.

Q to Dr. Hansen: Is there a correlation between overexpression of RANK and 18q?

A: He does not think so. All of the tumors have had this expression pattern, but he does not think this relationship is mandatory.

Cytokine/Death Receptor Pathways in Osteosarcoma

Fas Pathway - Eugenie S. Kleinerman, M.D.

Dr. Kleinerman began by stating that the morning’s discussion reviewed why osteosarcoma develops, but there is a difference between why it arises and why it metastasized. As one looks at tumors, tumor metastases have a propensity to grow in certain areas (i.e., osteosarcoma tends to grow in the lung).

In murine models, SAOS-LM6 cells were selected to have a more metastatic phenotype. Lung metastases with different propensity to metastasize to the lung were harvested. Fas expression was examined to determine its role in osteosarcoma metastases. Using Northern analysis, it was found that LM6 expressed a lower level of Fas. Flow cytometry showed that LM6 expressed a lower level of surface Fas.

To test whether Fas affects metastatic potential of the cells, Fas-transfected cells and control cells were injected into mice. The mice injected with control cells developed metastases, but the mice with Fas-transfected cells did not. Therefore, Fas may be very important to metastatic potential of osteosarcoma.

She then investigated whether Fas could be altered to impact metastatic potential. The adenovirus vector Ad.mIL-12 was used to transfect osteosarcoma cells to express IL-12. Cells that had been transfected had a reduced metastatic potential. IL-12 had an impact on Fas expression. IL-12 is composed of p40 and p35; it is only able to increase Fas expression when both of these subunits were available. Therefore, Dr. Kleinerman hypothesized that high Fas expression leads to tumor cell death and low Fas expression leads to increased growth.

Cells can also be transfected with the adenoviral vector through aerosol therapy. A study was conducted in which mice were injected with tumor cells, then aerosolized IL-12 adenovirus was administered. In the majority of mice receiving the adenovirus, pulmonary tumor cells were eliminated. This aerosol approach using liposomal 9-nitrocamptothecin is currently being tested in a Phase I trial for patients 10 years and older at MD Anderson.

Dr. Kleinerman also discussed INT-0133-Phase III Trial of Doxorubicin, Cisplatin and Methotrexate with and without Ifosfamide, with and without Muramyl Tripeptide Phosphatidylethanolamine (MTP-PE) for Treatment of Osteogenic Sarcoma. She considered whether Fas showed any relevance to this study. MTP can induce IL-12 in patients, because it activates macrophages which produce IL-12. 4-HC, the active metabolite of cyclophosphamide, upregulated Fas ligand expression in LM6 and LM6-IL-12 transfected osteosarcoma cells. The osteosarcoma cells were more sensitive in the presence of IL-12. She hypothesized that ifosfamide and IL-12 upregulate FasL and Fas respectively and induce osteosarcoma cell death.

Q/A Session

Q to Dr. Kleinerman: Could she predict the effects of ifosfamide alone?

A: There is data that shows if you upregulate Fas ligand, you can impair the immune system because CTL cells have Fas and thus the FasL on the tumor cell induces CTL death and immunosuppression.

Q to Dr. Kleinerman: What is the effector cell population in the nude mouse model?

A: The lung epithelial cells express Fas ligand. T cells don’t look like they play a role, but an increase in the influx of NK cells was seen in the nude mice that received IL-12.

Cytokine Apoptosis & Inhibition of Angiogenesis in Murine Solid Tumor Models - Jon M. Wigginton, M.D.

Dr. Wigginton discussed cytokine apoptosis and inhibition of angiogenesis in murine solid tumor models.

His rationale for expecting complementary antitumor effects by IL-12 and IL-2 is as follows: different signaling pathways are used by IL-12 vs. IL-2; IL-2 upregulates IL-12 receptor expression on T and NK cells; and the combination has enhanced immune stimulating effects for T and/or NK cells (cytotoxicity, cytokine production, and proliferation).

Evaluation has been done mostly in mouse-based renal cell carcinoma models, so the data presented focused on this tumor type. A pulse regimen which administered IL-2 over a series of days was conducted. The pulse regimen seemed to be well-tolerated and achieved tumor regression. Systemic IL-12/Pulse IL-2 inhibits tumor neovascularization and induces tumor regression via mechanisms which are dependent on CD8+ T cells, IFN-gamma, and Fas/Fas-L. The following statements can be made from the data:

  1. most mice cured of the renal cancer are resistant to rechallenge with the tumor
  2. the regimen is ineffective in SCID mice
  3. the tumor becomes infiltrated with CD8+ T cells, and depletion of these effector cells ablates the anti-tumor response

The role of IFN-gamma in the antitumor activity of this regimen is critical. Dr. Wigginton investigated whether the noted antiangiogenic effect played an early role in tumor regression. A recent study examined tumor angiogenesis samples 5 and 12 days after treatment ended. A trend of reduced vascularity to tumors was seen as early as 5 days of treatment. Therefore, IFN-gamma appears to play a central role in the significant reduction in tumor vascularization .

The impacts of IL-12/IL-2 on Fas/Fas-L gene expression may be as follows: RENCA (renal cell carcinoma) does not express Fas-L; IL-12/IL-2 potently enhances the production of IFN-gamma and TNF-alpha; and IFN-gamma/TNF-alpha enhances Fas expression on RENCA and EOMA (hemangioendothelioma).

The working hypothesis is that administration of IL-12/Pulse IL-2 induces Fas expression on tumor or endothelial cells which leads to apoptosis. The role of Fas in vascular endothelial injury induced by treatment with IL-12/Pulse IL-2 appears to be Fas ligand dependent, and antitumor activity against metastatic RENCA appears to be Fas ligand dependent.

A Phase I study of the IL-12/Pulse IL-2 schedule in adults has begun.

Fas & TRAIL-induced cell death in ES & NB – Maria Tsokos, M.D.

Dr. Tsokos discussed Fas and TRAIL-induced cell death in Ewing’s Sarcoma and neuroblastoma models. Fas has been implicated in a variety of tumors. If Fas is expressed on the tumor cell surface then it will die, unless it has developed a mechanism to prevent Fas triggered cell death. This could be done by downregulatation of Fas, Fas ligand, or Pro-Caspase 8 (FLICE), or upregulation of apoptosis inhibitory proteins, specifically bcl-2, or FLICE-inhibitory protein (FLIP). The Fas “counterattack” model refers to the Fas ligand expressed on tumor cell surface and its use to avoid activated T-cells, allowing tumor cells to escape surveillance.

Dr. Tsokos investigated the reason for Fas upregulation and how this occurs. Fas ligand exists in transmembrane and soluble forms in ESFT. She hypothesizes that ESFT cells may avoid cellular suicide by downregulating their surface FasL (by cleavage) and Fas (by internalization). When cells were treated with synthetic MMPIs, there were increased levels of surface FasL and Fas (in the absence of increased transcription), and cytoplasmic Fas is decreased. Exposure to MMPIs sensitized ESFT cells to a Fas-activating antibody (CH11) and sensitized cells to doxorubicin. Another study using TRAIL (an apoptosis-inducing ligand) was conducted. Receptor 5 (chemokine receptor 5, or CCR5) is needed to allow TRAIL to work and induce apoptosis. Lines lacking this receptor were the ones that were unable to respond.

Caspase 8 activation was found in Ewing’s Sarcoma cell lines, but not in the neuroblastoma cell lines. The lack of measurable expression of Caspase 8 in neuroblastoma does not translate to a lack of Caspase 8 presence, because FLIP is expressed in neuroblastoma. Neuroblastoma cells do not release cytochrome c from mitochondria during Fas activation, in contrast to ESFT cells. Caspase 3 is activated with antisense bcl-2; and Caspase 8 is important to antisense bcl-2 function (treatment of neuroblastoma with antisense bcl-2 caused cell death but only where Caspase 8 was present).

N-myc induces transcriptional upregulation of FasL, and in the presence of a functional Fas pathway, sensitizes neuroblastoma cells to apoptosis.

Interferon Gamma - Janet Houghton, Ph.D.

Dr. Houghton began by discussing the preclinical and clinical development behind modulation of the Fas signaling pathway in therapy for colon cancer. Fas has the following properties:

  1. type I transmembrane protein
  2. regulates apoptosis in cells for immune system
  3. expressed in many different cells and tissues
  4. expressed in colonic epithelial cells
  5. functional in cells outside of immune system
  6. Fas ligation induces apoptosis

Her studies confirmed that agents such as doxorubicin, topotecan, and VP-16 induced cytotoxicity without involvement from Fas. In observing Caspase 3 and Caspase 8 activity, she realized that Fas expression best correlated to apoptosis. Certain cytokines can elevate the level of Fas expression (specifically IFN-gamma). In a study investigating Fas expression in HT29 colon carcinoma cells after 2-hour exposure to IFN-gamma, the level of Fas was elevated, implying that IFN-gamma could sensitize cells to apoptosis. Potentiation of apoptosis by IFN-gamma is dependent on fluorouracil-induced DNA damage and the Fas death receptor, but is independent of p53.

The design of a Phase I clinical trial was to determine the maximum tolerated dose of IFN-gamma added to a standard regimen of 5-FU and leukovorin for metastatic GI cancers (19 patients enrolled; all colorectal cancer patients had prior exposure to 5-FU/LV regimens). Patient outcome data showed that 5FU disposition is not altered by IFN-gamma administration and there is a delayed absorption of IFN-gamma.

IFN-gamma upregulates Fas expression in HT29 cells at pharmacologically relevant concentrations, as determined in the plasma of patients treated in the phase 1 study. Exposure to IFN-gamma for 1 minute to 2 hours sensitizes HT29 cells to CH11 in clonogeneic assay. Finally, 5FU/LV combined with IFN-gamma induces cytotoxicity in HT29 at pharmacologically relevant concentrations. A Phase II trial is being planned.

Discussion - Holcombe E. Grier, M.D.

Q to Dr. Wigginton: How is, and how much is, Wyeth supplying IL-12?

A: They have an approved LOI for study in neuroblastoma in IL-12. IL-12 is unlikely to go forward as a single agent, and the drug’s future is dependent on results of the current clinical trials. There will be a company decision in about a year or so to see if cancer applications of IL-12 will continue.

Q to Dr. Hougton: She was asked about the great variability in the biology of osteosarcoma vs. colon cancer, and the fact that the data in the past panel was not directly related to osteosarcoma. Can this really be translated to osteosarcoma? Is there enough of a biological link between the previous presentations and osteosarcoma?

A: She responded that models of all sorts are needed.

A: Dr. Tsokos added that what is missing in all models is investigation of different pathways. We have great models, but don’t seem to generalize beyond a single tumor line.

Q: Dr. Gorlick asked what missing background is needed?

A: Dr. Tsokos responded that more cell lines and more pathways are needed.

Q: Any data on 5- FU and LV in osteosarcoma?

A: Data is from the 1970s saying it is inactive in osteosarcoma.

Q: It was mentioned that there are missing background studies. What background studies are needed?

A: It would be beneficial to have several cell lines that show response, and among the cell lines with no response, to determine why there is a lack of response by focusing on the upregulation or downregulation of certain proteins; very few people are doing lab research in osteosarcoma because of lack of funding and lack of interest in sarcomas.

Q: How would you look at the Intergroup trial in relation to the "counterattack" model? Is the B-minus result a significant result?

A: It could be real because you see the same results in two different patient populations, the non-metastatic and metastatic populations.

Q: We heard the hypothesis that MTP functions through IL-12…then IL-12 models and using knockouts of INF-gamma and effector knockouts…the question is, should we pursue immunologically-based trials in osteosarcoma? What we hear is to put together a hypothesis in a group of models to make it logically worth pursuing. Should we look at IL-12 if that is the major mediator? Should we do it in many different models to effectively dissect it?

Q: What’s the hypothesis?

A: Effects of IL-12, effected Fas, Fas ligand, and apoptosis. But now this needs to be further dissected and pursued in backgrounds where you knock-out various components to make it all fit.

Q: Any idea if GMCSF can perform similarly?

A: Aerosol GMCSF has been looked at in melanoma.

Q: What is the theory behind it being IFN-gamma independent?

A: You can upregulate Fas but not measure any IFN-gamma; this shows it is independent.

Q: Seems like IL-12 is fairly difficult to examine, but IFN-gamma is easier to look at. Any way to look only at IFN-gamma to see if that alone can do IFN-gamma?

A: There has been one trial with IFN-gamma (a Phase III ovarian cancer trial) that showed an advantage. Also, IFN-gamma is commercially available. But you need a model that is immunocompetent. The dog model could answer the GMCSF, IFN-gamma, and possibly IL-12 questions.

Drug Resistance/Growth Factor Pathways in Osteosarcoma

p-Glycoprotein (PGP) - Mark C. Gebhardt, M.D.

Although dramatic improvement in osteosarcoma patient survival has been seen with adjuvant chemotherapy, 20-40% of patients ultimately have tumor relapse, probably due to drug resistance. P-glycoprotein has been investigated as a means to identify these "non-responders" early, allow treatment modification, and potentially improve their survival. Similarly, PGP status may identify patients who could receive less toxic therapy.

A study was conducted in which PGP levels were viewed as a predictor of poor outcome in osteosarcoma. A variety of statistical tests determined that PGP+ patients were more likely to develop metastases and die (true for stage II B patients). Histologic necrosis was also investigated. Many other studies on PGP can be found in the literature, including Takeshita 1996, Baldini 1992, Wunder 1993, Baldini 1995, Baldini 1999, Stein 1993, Posl, 1997, Chan 1997, and Bodey 1995. Dr. Gebhardt described several possible optimal methods to assess PGP, which include: immunohistochemistry (but which antibody?); mRNA (RT-PCR), and in situ hybridization.

The possible mechanisms of drug resistance:

  1. multidrug Resistance Protein (MRP)
  2. topoisomerase II
  3. glutathione S-transferases
  4. enhanced DNA repair
  5. altered drug metabolism or inactivation
  6. reduced intracellular influx

MRP-positive cells have been detected in osteosarcoma. Caveats of the MDR theory in osteosarcoma include: the demonstration of a lesser survival in MDR-negative tumors may not be a reflection of decreased drug accumulation in PGP-positive tumors, the presence of PGP does not necessarily confer resistance, and its absence does not imply sensitivity.

The current literature concludes that there is no correlation between PGP status and percentage of necrosis. PGP status may predict disease outcome, but not histologic necrosis. The literature also concludes that PGP may be a sign of aggressiveness as well as a marker of drug resistance, and may be useful in identifying high-risk osteosarcoma.

The conclusions of Dr. Gebhardt’s study included:

  1. PPG expression is associated with a poorer metastasis-free survival and overall survival
  2. expression of PGP and its level were significantly predictive of time to death
  3. the monthly risk of death was four times higher among patients with PGP expressing tumors as compared to those with PGP negative tumors

A larger study, along with confirming the best assay of PGP, is needed.

Future needs include: determining whether PGP status is a predictor of outcome in larger groups of patients; how best to detect drug resistance (PGP and other mechanisms); risk-based therapy; other mechanisms of drug resistance; and newer chemotherapy and reversal agents.

p-Glycoprotein (PGP) - Irene L. Andrulis, Ph.D.

Dr. Andrulis continued the discussion of p-glycoproteins.

The study of molecular alterations in osteosarcoma include: identification and characterization of alterations in tumors; determining the clinical significance (prognostic or predictive value) of the alteration; and designing new therapeutic agents based on the molecular alteration.

Dr. Andrulis presented the same questions as Dr. Gebhardt: how are the "non-responders" identified, can they be treated differently and improve their survival, and are there patients who need less (or less toxic) therapy?

Improvements in therapy could result from investigation in the following areas: MDR1 expression in osteosarcoma; intrinsic resistance determined in biopsy specimens; resection specimen and relapse specimen-acquired resistance; low-level of drug resistance and subsequent detection in RNA; and looking at MDR1 expression in different sites within the same case (to determine whether the level of expression remains the same in different tumor sites within the same patient).

Drug resistance can be intrinsic (in the absence of treatment) or acquired (after treatment with chemotherapeutic agents). In a pilot study of 15 patients, the expression of MDR1 was correlated with poorer overall survival. A larger prospective study was then done. MDR1 expression was examined using RT-PCR and then correlated with DFS. Between 1989-1994, 123 newly diagnosed patients with high-grade, non-metastatic extremity osteosarcomas all received neoadjuvant chemotherapy that included doxorubicin. Although it was found that tumor size was predictive of outcome, the degree of tumor necrosis was not significant in the multivariate analysis (although it was in the univariate analysis). Patients with high levels of MDR1 did poorly, but patients with low levels of MDR1 also did poorly. Therefore, no association between survival and MDR1 could be made.

Dr. Andrulis inquired as to why there is a lack of correlation with MDR1 expression level and DFS in the prospective survival. The reasons include: conflicting results with IHC due to sensitivity of technique, antibodies detecting other proteins, and expression of MDR2 in osteosarcoma; other mechanisms of drug resistance; and p53 and drug resistance. Many osteosarcoma specimens have very high levels of expression of MDR1 and MDR2.

p53 is a tumor suppressor gene that is interesting because wild-type p53 can inhibit MDR1. p53 gene mutations in osteosarcoma fall in the standard genetic region, and in looking at other exons (including exon 4 and 10), some non-missense mutations are seen. When progressive sets of tumor biopsies from the same patient were examined at diagnosis and definitive surgery, it was seen that if the biopsy had a p53 mutation, then the resection had a mutation. When correlating type of p53 mutation and MDR1 expression, it was found that those who had missense had higher MDR1 expression, and those who had non-missense had lower levels of MDR1.

Ongoing studies are examining genetic alterations associated with progression and microarray data to identify novel genes/pathways.

INT-0133 MDR/PGP Study (Additional Presentation) - Cindy Schwartz, M.D.

Dr. Schwartz discussed INT-0133-Phase III Trial of Doxorubicin, Cisplatin and Methotrexate with and without Ifosfamide, with and without Muramyl Tripeptide Phosphatidyl Ethanolamine (MTP-PE) for Treatment of Osteogenic Sarcoma.

The objectives of the study were to determine whether multiple drug resistance encoded by PGP is useful to determine prognosis or to assign therapy, and to determine whether histologic response correlates with PGP positivity. The rationales behind the study included that PGP is an ATP-dependent efflux pump for hydrophobic substances and the presence of PGP has been associated with poor outcome in retrospective studies of patients with osteosarcoma. This study was intended to be the first large-scale prospective study of the effect of PGP expression on outcome. 679 patients with non-metastatic disease from CCG and POG were enrolled on this trial between November 1993 and November 1997. Antibodies used included C-494 and JSB-1. Results showed that 107 and 104 patients were evaluated with C-494 and JSB-1, respectively, and discordance was only 14% between C-494 and JSB1. EFS, osteosarcoma, and histological response were identical in those evaluated vs. those not evaluated. EFS and increased risk of death did not differ significantly in patients with tumors that demonstrated antibody positivity at study entry compared to antibody negative patients. The patients on regimen A (no ifosfamide) with antibody positivity at study entry were not at significantly increased risk for adverse events or death. The patients on Regimen B (+ ifosfamide) with antibody positivity at study entry were not associated with increased risk of adverse events, but there was evidence of a reduced risk of death in these patients. Each regimen had a chemotherapeutic agent with a PGP substrate in them.

The study concludes that MDR-encoded PGP is not useful to determine prognosis in patients with localized osteogenic sarcoma treated with cisplatin, methotrexate, doxorubicin, + MTP-PE, + ifosfamide. Furthermore, histologic response is not correlated with PGP positivity. It is unclear whether further evaluation is needed to determine the usefulness of MDR-encoded PGP to assign therapy.

Discussion - Leonard H. Wexler, M.D.

Q: Given the seeming discrepancy of results that show no difference in the percentage of necrosis but does demonstrate a difference in survival, should this be looked at in 500 patients? If you do decide to look at it, how do you look at it?

A: Most of the studies have done it on biopsy materials, and not put the tumor into culture. A lot of care should go into a proper design.

Q: So is MDR an area that the groups should look at?

A: No consensus of the group was reached on this topic.

Q: What is the status of MDR reversal agents in other tumors?

A: Dr. Smith responded that overall results have been disappointing, and there have been few successes. This should not be a high priority, but it can be looked at some. Other participants said that this should not be investigated much further.

A: If doxorubicin is the major agent, then there may be some merit to it. In the Toronto study, RT-PCR and IHC were done and very few were positive.

Dr. Andrulis commented that PGP does not have to be constantly measured, but wouldn’t discount this yet. In their study patients with very, very low MDR1 as well as those with very high levels of expression were those patients with a poor prognosis. Microarrays should be used to see what is happening in these pathways.

Q: Should we continue to do MDR related studies?

A. No consensus of the group was reached on this topic.

Antibody Based Strategies for Metastatic Solid Tumors - Nai-Kong V. Cheung, M.D., Ph.D.

Dr. Cheung discussed potential therapy targets using mouse or chimeric antibodies, anti-idiotypes as vaccines, pretargeting strategies, and retargeted T-cells.

Potential targets are GD2 (monoclonal antibodies 3F8 and 5F11) and gp58 (antibody 8H9). 3F8 targets GD2 (disialoganglioside) on neuroblastoma. GD2 is abundant on neuroblastoma (5-10x106) and shows persistence on the cell surface. gp58 is seen in many pediatric tumors, including osteosarcoma.

Mouse or chimeric antibodies constitute active therapy alone or with BRM (Beta-glucan). Beta-glucan enhances the anti-tumor effect of monoclonal antibodies. Used alone, Beta-glucan has no anti-tumor effect, but with MoAb the effect is significant and in mouse tumor models, Beta-glucan + MoAb prolongs progression-free survival.

A pretargeting strategy uses scFv-streptavidin (radioisotopes, biologics). Following tumor targeting of scFv-streptavidin, a clearing agent such as biotin-LC-NM-(Gal-Nac)16 removes residual fusion proteins from the blood, significantly improving the tumor:blood AUC ratio. The whole antibody has good uptake but not a good ratio, while fragments give a better ratio but poor uptake. Using GD2 as a target, whole antibody or antibody-fragment alone results in a tumor:blood AUC ratio of <5:1 , while the pretargeting approach gives a >100:1 ratio. This system allows a high-level of uptake in the tumor and is very robust.

Retargeting T-cells involves modification of T-cells and development of chimeric immune receptors (T-bodies). All T-cells would recognize the same antigen and would proliferate when the anti-idiotype (8H9 scFv-CD28-symbol chimeric immune receptor) is added. When used in rhabdomyosarcoma, tumor suppression is seen.

Growth Hormone - Lee J. Helman, M.D.

Dr. Helman discussed Insulin-like Growth Factor -1 (IGF-1) and osteosarcoma.

Characteristics described for IGF-1 and osteosarcoma are as follows: peak incidence of osteosarcoma occurs during peak GH and IGF-1 concentrations; prevalence of osteosarcoma is highest in dog breeds with the highest IGF-1 concentrations; hypophysectomy is associated with anti-tumor activity in rodent osteosarcoma models (anti-tumor effect is associated with IGF-1 inhibition); IGF-1 inhibition by growth hormone blockade results in decreased osteosarcoma growth; and osteosarcoma cell lines are dependent on IGF-1 for in vitro growth and survival.

Osteosarcoma cells express IGF-1 receptors, are stimulated to proliferate by IGF-1, and are more chemoresistant after IGF-1 addition. There is interest in taking advantage of this because there are many inhibitors available for this pathway.

SMS 201-995pa LAR (OncoLar - Sustained Release Sandostatin) is a long acting formulation of octreotide acetate. Octreotide acetate can effectively inhibit IGF-1 through the blockade of growth hormone.

A Phase I study (95-C-0119) of OncoLar with/without Tamoxifen in relapsed osteosarcoma administered 60 mg to 6 patients, 90 mg to 6 patients, 60mg plus tamoxifen to 6 patients, and 90 mg plus tamoxifen to 3 patients. OncoLar was administered every 4 weeks, and growth hormones were measured; the only sustained decrease was in circulating IGF-1. This was not dose-dependent and addition of tamoxifen had no influence.

The study concluded that OncoLar treatment leads to a sustained decrease in IGF-1 levels (40-50%) that was dose independent between 60 mg and 90 mg. The study also concluded that no other component of the IGF pathway is altered, and there was no additional decrease with tamoxifen. At the level of IGF-1 reduction achieved, however, no measurable tumor response occurred. growth hormone receptor

Future studies to consider include:

  1. combination with cytotoxic therapy
  2. development of a Pegvisomant- growth hormone receptor analogue that prevents growth hormone receptor dimerization and is a stronger inhibitor of circulating IGF-1
  3. combinations of growth hormone receptor antagonists with Somatostatin analogues

Dr. Helman posed two hypotheses. First is that the inhibition of the IGF-1 growth signal, combined with DNA damage from conventional chemotherapy, may act in an additive or potentially synergistic manner to induce tumor apoptosis. Also, enhanced tumor apoptosis may significantly improve survival outcomes in patients with osteosarcoma without increasing treatment related toxicity.

Q/A Session

Q: Would a 50% reduction in IGF-1 be expected to cause significant anti-tumor activity?

A: No.

IGF - Jeffrey A. Toretsky, M.D.

Dr. Toretsky asked whether IGF-1 contributes to tumorigenesis, tumor growth and progression, or chemotherapy resistance; and can modulation of IGF-1 reduce tumorigenesis, reduce tumor growth and progression, or modulate chemotherapy resistance? Epidemiologic considerations include: 1) does growth predispose to osteosarcoma; 2) is increased size/rate of growth adequate to implicate IGF-1 level as a risk factor; 3) and do elevated levels of IGF-1 predispose to osteosarcoma? Multiple studies have addressed whether increased growth (both weight and height) identify persons more likely to develop osteosarcoma. The three most recent studies are case-controlled, however all utilized different control groups. When either a normalized growth curve from National Center for Health Statistics or geographic neighbors were used, no difference in final height or growth velocity were observed between cases and controls. The most recent study utilized controls matched from birth data and evaluated the height and weight one year prior to diagnosis. This study did show that size one year prior to diagnosis was a risk factor for developing osteosarcoma. The authors did suggest selection bias might play a role in their results. I believe overall it is unclear whether "big" people are more likely to develop osteosarcoma.

IGF-1 activity is activated through its cognate transmembrane receptor; which can also be activated by IGF-2. The IGF-IR is required by many oncogenes for transformation (IGF-1R null mouse embryonic fibroblasts do not transform independently with SV40 lg T, Ha or Ki ras, or c-myc); and the signaling pathway activated by IGF-1R leads to cell survival.

IGF-1R studies in osteosarcoma have determined the following: human tumors express IGF-1R; cell line signaling/cell survival is inhibited by aIR3-antisense, or suramin; and p53, occasionally mutated in osteosarcoma regulates IGF-1R expression.

A recent study by Rodriguez-Galindo, et al, 2001, determined that circulating concentrations of IGF-1 and IGFBP-3 are not predictive of incidence or clinical behavior of osteosarcoma. However, the study was limited by the small number of patients and lack of statistical power. Further investigation regarding the role of the IGF-1 pathway in the development and progression of osteosarcoma is warranted, and suggested by the authors, and could conceivably provide new potential targets for therapeutic intervention.

Dr. Toretsky investigated the relationship of IGF-1 levels to EFS in the Ewing’s Sarcoma Family of Tumors (ESFT). Using data from 111 patient samples available for analysis, it was shown that serum IGF-1 and IGFBP-3 proteins survive prolonged freezing. Preliminary studies also showed that IGF-1 and IGFBP-3 levels in ESFT patients can identify patients with the most widespread disease, but are not independent predictors of prognosis. A high IGFBP-3 to IGF-1 ratio suggested a better prognosis. Improved survival among metastatic patients was also associated with IGFBP-3 to IGF-1 ratios that fell between the 75th and 90th percentile of the samples measured. This "Mama Bear Effect" is seen in those patients with a mid-range level of free IGF-1 that ultimately show the highest survival rate. Patients with very high ratios had extremely low IGF-I levels and had severe systemic disease. Those with low ratios might have had high levels of IGF-I and thus potentially did poorly in part due to IGF-1R modulated chemoresistance in a small population of cells that ultimately regrow. While there is in vitro data to support this hypothesis, a prospective trial is underway to determine if this result is reproducible in patients currently undergoing therapy.

Current plans for osteosarcoma include: investigating collected serum from 214 patients as part of protocols (P9851); measuring IGF-1 and IGFBP-3; and evaluating these results as a function of DFS in the next several years. We are considering measuring receptor levels as well, however, it is challenging to attempt quantitation of IGF-1R utilizing either immunohistochemistry or ELISA. There is a potential for new therapies which target block ligand binding sites on the receptor, inhibit the receptor kinase activity, or by targeting downstream signaling.

Dr. Toretsky concluded that IGF is likely a part of a multifactorial pathogenesis of osteosarcoma and needs to be better characterized in a wider panel of osteosarcoma cell lines and patient samples. Current studies will identify whether IGF-1 levels play a role in chemosensitivity/relapse. Inhibition of IGF signaling may provide a future therapy against osteosarcoma, but more background and better agents are needed.

Discussion - Neyssa Marina, M.D.

Q: What are IGF-1 levels in normal patients? How do your results relate to normal?

A: That is very difficult to get to because IGF-1 normal values have a large range which varies with age. It is better to look at it in regards to puberty, but the data just isn’t there. There isn’t a good control group.

Q: Is there any evidence that the receptor is abnormal? Maybe not abnormal levels, but rather during growth a receptor malfunctioned and doesn’t know how to shut off signal?

A: There is not much information suggesting a mutated receptor. Currently there are no known activation mutations occurring in patients. No one that I am aware of has done systemic searches into receptor mutation for the IGF-I receptor. With regard to the epidemiology, the thinking has changed over time. There are epidemiological data in epithelial cancers that says very high IGF are associated with a higher risk. But a predominance of evidence says that IGF has some role. It is a necessary but possibly insufficient for tumorigenesis. Once a tumor is there, components of pathways can be blocked and cells can be pushed to make a death decision.

Q: Are there pharmaceutical companies looking into receptor inhibitors for osteosarcoma?

A (Dr. Helman): As of last May, four companies. When you inhibit IGF-1, insulin is inhibited which is obviously not a good thing to do. This must be targeted without harming the insulin receptor. These unique compounds should be available in the next few years.

A (Dr Toretsky): There are compounds being developed but people are very tight-lipped about it.

Q: Do cells still die with dominant negative AKT? The survival curve is different, the dominant AKT cells die at same point that the other cells don’t. Can you rescue the AKT cells?

A: This hasn’t been researched yet.

Q: Can you utilize the fact that osteosarcoma is right outside the growth curve? And possibly see whether or not there is a difference between a young patient and a patient in the middle of a growth spurt?

A: A study has already been done on this, and didn’t see a difference between these two groups. No cytogenetics were available; it was strictly a RNA/DNA study.

Q: It seems that for a lot of these agents they have to be combined with chemotherapy. What preclinical data do you need to see and what data can we gather for 3F8 to convince you to go ahead with a Phase III trial since you may never get a Phase II trial?

A: There are several. But compelling animal data (specifically in dogs) that shows blockage of pathways is key. A caveat is that the route would have to be relatively non-toxic since it may be needed for a lengthy period of time.

Q: Maybe look at rhabdomyosarcoma situations. There is a dwindling but still sizable number of patients with osteosarcoma who relapse. Maybe these patients could be used to answer some of these questions, and give them a biologic agent to show efficacy. There may be ethical issues which could be covered with crossover design.

A: A possible trial would be patients with A plus A+B, but this would be a Phase III question because it may be more toxic and costly and in need of more power. There should be two things at the same point in development and compare them in Phase II to see which one moves forward to Phase III. An example is aerosol GMCSF vs. an AKT inhibitor.

Q: How to prioritize? Which drugs are most important to move forward? Should there be excitement about the antibodies mentioned by Dr. Cheung?

A: It is a reasonable target and an imaging trial will be starting soon which will give some more information. No imaging studies are planned with the GD2 antibody, but it is a reasonable system. All sarcomas are eligible for upcoming trial. As for the anti-idiotypic antibody for GD2, a study has been on for 4-5 years with slow accrual. Some antibody responses are being seen, but it is early.

New Treatment Approaches

Isotopes: Samarium - Peter M. Anderson, M.D., Ph.D.

Dr. Anderson presented data regarding Samarium (153Sm-EDTMP, ethylene-diamine-tetra-methylene-phosphonate), which is a bone-seeking radiopharmaceutical that has been FDA-approved since April 1997. Samarium emits a short-range therapeutic beta particle and also has a gamma photon that allows routine gamma camera imaging. It has a short half-life of 47 hours. The dose-limiting toxicity has been hematologic toxicity because it binds to bone causing the tissue to become radioactive. Thus, bone marrow is also affected. However, the short-half life of samarium allows for rescue of the marrow with stem cells.

A study conducted at the Mayo Clinic administered high-dose Samarium to patients with bone metastases (JCO 2002,20:189-196). Accrual was difficult and the study was expanded to include myeloma patients. Patients were all high-risk and ranged in age from 12-43 years, with a mean age of 20.5. To date, the results show: flat dosing; no "scout dose"; and, if the dose is > 3 mCi/kg, the patient is required to get peripheral blood stem cell (PBSC) rescue. The uptake into the bone is linear with dose. A technetium-avid bone scan predicts tumor uptake of the samarium. "Bone specific" palliative RT for resistant disease was permitted. All patients have neutropenia and require transfusions and GMCSF.

Administration is by 30-minute central line infusion. Dosimetry at Day +1, +2, and +5 allows estimates of residual radioactivity at time of PBSC infusion and dose to tumor. The Day 13 radioactivity estimate was <3.6 mCi. Infusion of PBSC was done on Day +14. In terms of toxicity, there was minimal non-hematologic toxicity, hypocalcemia, and hematologic toxicity. Hypocalcemia was manageable, but the hematologic toxicity resulted in a poor recovery if there was a low number of PBSC (<2 x 106 CD34+/kg). Future studies should not give high-dose samarium if there is inadequate PBSC supply stored. A "flare" reaction of bone pain was common, but improved pain control was seen in 100% of patients. Osteosarcoma recurrence (both bone and lung) remains a major problem after Samarium treatment in this high-risk population of patients presenting with bone metastases.

In conclusion, Dr. Anderson stated that samarium is a bone-specific therapy that may provide more effective radiotherapy than external beam radiation alone. It appears safe, effective, and well-tolerated, except for temporary pancytopenia. A future clinical trial concept would be to compare high-dose vs. standard dose samarium (if there are no PBSC for harvesting) with the primary endpoint of radiation dose to indicator lesion and secondary endpoints of PET responses, time to progression, pattern of disease progression, and ability to provide additional therapy. Use of a radiation sensitizer (low dose gemcitabine) for several days after the samarium may be another means to increase radiobiologic effectiveness.

Q: Was chemotherapy administered to any patients after Samarium?

A: Yes, some patients received ifosfamide/etoposide and some received Gemcitabine. This depended on prior therapy and dose of stem cells (generally complete recovery if >5 x 106 CD34+/kg)

Inhaled GMCSF - Carola A.S. Arndt, M.D.

Dr. Arndt discussed aerosolized GM-CSF for patients with osteosarcoma and lung metastases. She presented background information from Dranoff (1993) who studied GM-CSF in the B-16 melanoma model. The B-16 melanoma cells were transduced with various immunomodulatory proteins; mice were vaccinated with irradiated non-transduced or transduced cells, and challenged with live non-transduced B-16 cells. Irradiated tumor cells expressing GM-CSF reliably stimulated potent, specific, long-lasting immunity in three different tumor models; GM-CSF was the best of >10 cytokines tested. GM-CSF resulted not only in anti-tumor response, but also protected mice from subsequent tumor challenge.

Dr. Arndt also described work by Kumar (1999) who studied the localized effects of GM-CSF and found that GM-CSF transduced melanoma cells had lower tumorigenicity than control cells. Tumors transduced with GM-CSF had dense macrophage infiltration, and local GM-CSF produced by transduced cells facilitated killing of non-transduced bystander tumor cells (co-culture experiments). GM-CSF also enhances marrow recovery post-chemotherapy, modulates the host defense versus bacterial fungi, and acts as a vaccine.

One of the principles of aerosol delivery for inhaled GM-CSF is that small particles take a long time to settle. Nebulizers provide >95% particles of 0.4-3.0 u, of which about 15-20% are deposited in lungs. Aerosol cytokines have the following characteristics: lymphatic absorption; receptor bearing cells in lymphatics, bronchial, and pulmonary lymph nodes; cells provide receptor bearing "gauntlet"; and the effect is dose dependent.

The Phase I dose escalation study (Mayo 97-02-01) allowed intra-patient dose escalation if the patient tolerated the initial dose well. There were two osteosarcoma patients out of seven patients that completed all three dose levels. Of the two, one patient had "too numerous to count" metastases and progressed, and one patient was stable for 11 months. Five of six patients had stable disease and continued aerosolized GM-CSF for an additional two to eight months. No toxicity was seen in any patient.

A Phase II study with 45 patients was completed with eight osteosarcoma patients. Two of eight patients with gross disease progressed. The remaining six achieved second CR, with two patients disease-free at 2.5 and four years.

Dr. Arndt’s proposal for a follow-up study begins with the administration of aerosolized GM-CSF to patients with isolated pulmonary recurrences of resectable osteosarcoma. A thoracotomy will be performed one week later, and a descriptive analysis of histologic findings will be conducted. The time to next recurrence will be analyzed, and patients with bilateral metastases will have initial thoracotomy. Upon recovery, inhaled GM-CSF will be administered for one week, followed by second thoracotomy and continued GM-CSF. Histologic evaluation would include: nodule evaluation for evidence of expression of Fas/FasL; nodules evaluation by routine H & E for evidence of macrophage and neutrophil infiltration; and nodule evaluation for the presence of dendritic cells by immunostains. A similar retrospective tissue study of resected lung metastases at first recurrence prior to chemotherapy is being planned to obtain baseline information.

In the future, Dr. Arndt proposes a pilot study testing administration of inhaled GM-CSF to patients with localized, non-metastatic osteosarcoma beginning right after surgery and continuing for six months. She also proposed inhaled GM-CSF use for patients with pulmonary metastases at diagnosis, post-surgery for lung metastases, and post-therapy in patients at high-risk for future lung metastases.

Antifolate Resistance - Richard Gorlick, M.D.

Dr. Gorlick presented information on antifolate resistance, specifically looking at methotrexate. Most soft tissue sarcomas are intrinsically resistant to methotrexate. The literature suggests it is the peak methotrexate level, and not the AUC, that correlates with therapeutic response in osteosarcoma - the peak level is predominantly determined by dose. High-dose methotrexate is a major component of therapy for osteosarcoma, and tumor resistance to methotrexate is known to occur. The reasons for methotrexate resistance may include intrinsic or acquired methotrexate resistance. The use of neoadjuvant chemotherapy only allows the study of viable/resistant tumors after methotrexate has been administered and all at the time of initial biopsy. Examination of tumors also obtained from the 30% of patients who relapse can be used to study acquired methotrexate resistance.

Trimetrexate has been investigated as an alternative approach for antifolate therapy. A Phase II trial of simultaneous trimetrexate and leucovorin given orally for 21 days was conducted in relapsed osteosarcoma patients. Toxicity was acceptable with myelosuppression being the major side effect, and objective responses were observed in 8%, which may be good in this population (n=39, CR=1, PR=2, MR=1, SD=8).

A phase II window study proposal is under development by the COG with the rationale that the response rate to trimetrexate and simultaneous leucovorin would be predicted to be higher in the window setting than in a conventional phase II trial because dihydrofolate reductase overexpression is not as frequent at diagnosis as compared to relapse. Furthermore, trimetrexate with leucovorin would be predicted to be at least as efficacious as high-dose methotrexate with less toxicity. Also, the lack of renal toxicity may allow chemotherapy dose intensification in subsequent clinical trials, as it may be possible to administer trimetrexate simultaneously with other agents.

Viral Approaches - Thomas A. Gardner, M.D.

Dr. Gardner discussed an osteocalcin promoter-based cancer gene therapy for metastatic disease. There are three components of gene therapy: vectors (naked DNA; liposomal; virus: retroviral, adenoviral, AAV, etc.), transcriptional regulators (universal: LTR, CMV, RSV, SV40, etc.; tumor/tissue-specific: PSA, CEA, OC, AFP, etc.), and therapeutic genes (replacement: p53, CFTR, etc.; toxic: TK, CD, P450, DT, etc.).The positive attributes of using adenoviruses as a viral vector for gene therapy include: high transduction efficiency, large insert capability, non-integration into host DNA, and broad natural tropism. Negative attributes include: moderate immune response and broad natural tropism.

Osteocalcin is a non-collagenous Gla protein produced by osteoblasts. It is synthesized and secreted during bone mineralization, and its immunohistochemical staining is positive in osteoblastic osteosarcoma (100%), chondroblastic osteosarcoma (100%), and fibroblastic osteosarcoma (71%).

Recent osteocalcin promoter-based gene therapy studies have shown in vitro and in vivo growth inhibition of rat osteosarcoma with ad-OC-TK/ACV (adenoviral osteocalcin gene vector). These studies have also shown enhanced in vitro and in vivo growth inhibition of rat and human osteosarcoma with the addition of 5-FU chemotherapy to Ad-OC-TK/ACV and in vivo growth inhibition of metastatic rat osteosarcoma with systemic Ad-OC-TK/ACV. In vitro and in vivo growth inhibition of metastatic human prostate cancer with Ad-OC-TK/ACV has also be demonstrated in preclinical models.

A Phase I trial in prostate cancer (OBA Protocol #9812-276) evaluated safety, gene transfer, and activity of the gene therapy approach. This study showed that repeated intratumoral injections were well-tolerated. No viral shedding was observed after Day 3 and no unexpected toxicities occurred.

Dr. Gardner then discussed Ad-OC-TK / ACV gene therapy for osteosarcoma, the development of chemo-gene therapy for osteosarcoma, the potential suppression of osteosarcoma pulmonary metastases , and the use of isolated lung perfusion to enhance gene therapy delivery.

A new Phase I trial (OBA #0010-426 - A Phase I Study of Intratumoral Injections of OCaP1 into Osseous Metastatic or Post-Surgical Locally Recurrent Prostate Cancer) was reviewed and a possible Phase I Trial of Ad-OC-E1a Pulmonary Metastatic Osteosarcoma was discussed.

Dr. Gardner summarized his presentation as follows: Ad-OC-TK/ACV can inhibit ROS (reactive oxygen species) and MG-63 (p53 gene mutated) cell growth in vitro; intralesional Ad-OC-TK/ACV can inhibit ROS and MG-63 growth in vivo; chemotherapy can enhance cancer gene therapy; systemic Ad-OC-TK/ACV can inhibit ROS pulmonary metastases; isolated lung perfusion can concentrate adenoviral delivery to further inhibit pulmonary metastases; and Ad-OC-TK/ACV maintains potency in osteocalcin+ cell lines and can be administered systemically.

Q/A Session

Q: Does this localize to growing bone?

A: The answer is yes. We need to address this more, possibly in adolescent rodents.

Q: Any data on presence of osteocalcin levels in patients with osteosarcoma?

A: Yes, high levels are seen in a series of metastatic and primary osteosarcomas. There are also cases where they are not so typical.

Slow Release Platinum – Stephen J. Withrow, D.V.M.

Dr. Withrow discussed slow release platinum (open cell polylactic acid and cisplatin - OPLA Pt), and presented a dog model of osteosarcoma. OPLA Pt is administered via a sponge-like device that can be sized and shaped to necessary specifications. An immediate drug dose is given upon insertion, but the small holes of the sponge device allow for slower drug release. The sponge delivery device is biodegradable, like an absorbable suture.

The rationale for drug delivery by such a device includes: 1) the problem of local recurrence in many solid tumors; 2) some sarcoma patients are not salvaged because of marginal resections; 3) most marginal resections leave tumor close to the wound bed; and 4) intracavitary treatment increases the local dose while decreasing the systemic dose. Preliminary work includes in vitro elution in serum, mouse-transplantable carcinoma, and dogs-cortical allografts (tumor bearing dogs). If intravenous platinum is given, it is taken in quickly and depleted quickly. If the implant device is used, the release of platinum is slow.

The applications for OPLA include its use at the local tumor site for limb sparing-procedures or for soft tissue sarcomas. This delivery of platinum also acts as a radiation sensitizer.

A protocol design for canines was presented. Over a period of 10 years, 80 dogs had surgical resection of osteosarcoma with allograft placement. Half of the dogs received the platinum implant in the wound, and all dogs received I.V. chemotherapy. In the implant group there was a 15% recurrence rate, while the control group had a 55% recurrence rate. Among the dogs who had incomplete resection of their tumor, there was a statistically significant decrease in the tumor recurrence rate among the group that received the platinum implant .

Problems associated with the use of the implant can include infection and slow degradation of the polymer. Issues that remain unresolved include: 1) the best carrier and the best release profile; 2) the influence of a surgical drain and the radiation-heat effect on the drug activity; 3) whether this delivery technique will work with other drugs; and 4) whether this approach will be active in both local disease and metastatic disease.

Antiangiogenic Approaches - Mark W. Kieran, M.D.

Dr. Kieran described angiogenesis and the agents available to inhibit it. He then discussed potential surrogate markers of response, and reviewed issues of trial design. He defined anti-angiogenesis as the inhibition of the induction, stabilization or progression of neo-vascularization. The vascular endothelium, pericytes, vascular basement matrix, and bone marrow cells are all components of angiogenesis that can also be targeted. Regulatory molecules include: Ang1, Tie2, Ang2, VEGF, EphB2, and EphB4.

Available anti-angiogenesis agents include SU5416, squalamine, thalidomide, celecoxib, oral cytoxan, and oral VP-16. SU5416 is a small molecule inhibitor of VEGFR that is currently in a Phase I PBTC trial. Squalamine is a NHE3 pump inhibitor in a new Phase I trial in COG with carboplatin. Thalidomide is currently in phase II trials with other agents; it inhibits bFGF and VEGF signals. Celecoxib inhibits COX-2 activity of endothelial cells and is in Phase II trials in combination with other agents (including celecoxib, oral cytoxan and oral VP-16 above). There are a variety of oncogenes that relate to angiogenesis including K-ras, H-ras, c-myb, N-myc, c-myc, HER-2, EGFR, and PyMT for which inhibitors are now available. STI571 and ZD1839 are now being studied for their antiangiogenic effects.

There are currently no good surrogate markers of angiogenic response; even standard radiographic markers of response are inadequate for efficacy. Treatment endpoints including microvessel density, bFGF, VEGF, PDGF, etc levels, endothelial proliferation, migration or tube formation, bone marrow derived endothelial precursors are being studied as surrogate markers of angiogenesis inhibition.

The standard phase II trials identify a radiographic treatment response as a complete response, partial response, or stable disease. Progressive disease is defined as a greater than 25% increase in tumor size. In clinical trials of antiangiogenic agents, patients would remain on study unless the increase in tumor size was greater than 100% . Patients would undergo response evaluation at 8 weeks, and patients with progression (if clinically appropriate) could remain on therapy for up to 16 weeks. The reason for this stems from the over-abundance of tumor induced blood vessels. Unlike standard chemotherapy which shrinks the tumor within a few weeks if effective, antiangiogenic agents that kill part of the tumor vasculature can leave enough vessels behind so that the tumor can keep growing. Only after prolonged therapy (depending on the treatment and tumor) will tumor stabilization and eventually tumor regression be possible.

Issues to be considered in the design of antiangiogenic agent clinical study include the length of time a patient should remain on therapy and the timeframe within which a tumor response should be expected. In addition, if a patient demonstrates tumor progression on an antiangiogenic agent, one could add additional agents while continuing the initial one since endothelial cell resistance is unexpected. Tumors that initially respond and then progress on an antiangiogenic agent may have induced other stimulators of neovascularization. As such, adding in agents to cover the new stimulator may require that the original stimulator continue to be suppressed.

In summary, Dr. Kieran stated that a limited number of antiangiogenic agents are available in pediatrics. The issues related to study endpoints and the lack of good surrogate markers of response make evaluation of agent activity difficult. The potential application of new radiologic technologies is limited by expense.

Bisphosphonates - John H. Healey, M.D.

Dr. Healey discussed bisphosphonates, which were first identified in 1965. The structure is based on pyrophosphate, and there are two families of bisphosphonates. There are three essential elements to their activity: direct effects on mature osteoclasts; effects on osteoclast precursor cells; and effects on osteoblasts. The bisphosphonates are negatively charged and the means of cell entry is still unknown. These chemicals bind to the bone for the duration of the half-life of the bone, which is 10-12 years in humans, three years in dogs, and 200 days in rats.

The activities of bisphosphonates in bone cells include apoptosis, inhibiting cytokines, inhibiting metalloproteinases, disrupting adhesion molecules, immunomodulation against myeloma, and antiangiogenesis mechanisms. Specifically, bisphosphonates inhibit IL–6, a cytokine which promotes myeloma and osteosarcoma cell growth and encourages angiogenesis. Inhibition of metalloproteinases is important because metalloproteinases are present in bone resorption patients without osteoclasts.

Research in bisphosphonates and fracture healing has shown the following: etidronate causes ostomalacia and retards fracture healing; alendronate at 2 mg/kg/day caused larger, less mineralized callus that had normal strength; incadronate (YM-175) at 10–100 µg/kg for long-term (25-49 weeks) in rats delayed fracture healing, but it didn’t impair the mechanical strength.

Bisphosphonates prevent bone loss and fractures, and also have the ability to prevent osseous and non-osseous metastases. This is well established in patients with breast cancer and microscopic bone marrow "metastases" on staging studies. (Diel, IJ NEJM 98) Potential roles for bisphosphonates in sarcoma patients include preventing metastases, preventing osteoporosis, preventing periprosthetic osteolysis from particulate debris, and preventing bone loss from the stress bypass effect. It is clinically important to look at osteoporosis in children with sarcoma. Nine of 108 lower extremity osteosarcoma and Ewing's Sarcoma patients treated ‘90-’99 suffered a lower extremity fracture due to treatment-related osteoporosis.

In preclinical studies, Walker 256B carcinoma was implanted intraosseously into Wistar-Lewis rats to view experimental metastases. Pretreatment with clodronate inhibited the development of bone metastases compared with controls. Shorter intervals between the bisphosphonate therapy and the inoculation of tumor cells gave the best results. This data suggest, that in the human setting, early therapy might give better results.

Toxicities associated with bisphosphonates administration include fever, uveitis, leukopenia (transient, mild), and myalgia (GI upset rare). There have been no particular problems reported in children who have been treated for osteogenesis imperfecta, hypercalcemia of malignancy, and fibrous dysplasia.

In conclusion, bisphosphonates are active against osteoporosis, in prevention of metastatic tumor development, and possibly against primary tumor progression. These agents are safe in adults, but the data with children is limited.

Q/A Session

Q: Dr. Kieran was asked what should be done now to look at angiogenesis in osteosarcoma?

A: Right now thalidomide, Celebrex, and other agents are being used. A trial started two months ago at DFCC has already completed half of its accrual.

Guest Presentation: Safety and Pharmacokinetics of Celecoxib when Combined with Metronomic Administration of Low-Dose Chemotherapy of Pediatric Recurrent Solid Tumors - David Malkin, M.D., on behalf of Sylvain Baruchel

Dr. Malkin discussed the combination of metronomic chemotherapy and celecoxib as an antiangiogenic treatment. Metronomic dosing of chemotherapy uses lower doses administered more frequently, blocks endothelial cell proliferation, appears more tolerable, and there is no acquired drug resistance (endothelial cells are normal, homogeneous, and genetically stable). COX-2 is expressed within tumor neo-vasculature, neoplastic cells, and stroma cells (quiescent vasculature only expresses COX-1). Celecoxib is antiangiogenic at pharmacologic doses. Vinblastine is also being used.

COX-2 is upregulated in most human tumors. COX-2 -/- mice have decreased VEGF expression, reduced tumor angiogenesis, and tumor growth. COX-2 inhibitors decrease VEGF production and prevent VEGF-induced MAPK activation in the endothelial cell compartment. COX-2 inhibitor suppresses activation of small GTPase Cdc42/Rac and alphaVbeta3 integrin in vitro and in vivo.

The objectives of this study are to evaluate the safety and toxicity of chronically administered low-dose chemotherapy plus celecoxib. Pharmacokinetics are also examined, and a response rate is estimated. Other objectives are to validate angiogenic growth factors (VEGF, bFGF, VCAM-1) as surrogate markers of angiogenesis, and to evaluate dynamic MRI (dMRI) as a tool to assess specific antiangiogenic tumor response. The study design is open label, non-randomized, Canadian multicenter Phase I/II pilot study, for patients >21 with histologically or cytologically confirmed solid tumor (excluding brain stem tumors) measurable by radiographic imaging. dMRI parameters are being used to provide tissue perfusion, blood volume, capillary permeability, and extravascular volume.

Thus far, 17 patients have been enrolled, with five having osteosarcomas. Celecoxib has been well tolerated. There have only been three adverse events reported, although they are most likely due to cyclophosphamide and not celecoxib, except for a rash.

Surrogate markers of tumor angiogenesis have a large interpatient variability (especially with VEGF and bFGF), and more patients are needed to validate these biomarkers. However, VEGF and bFGF showed a trend of decreasing during the course of treatment despite observing tumor progression.

Challenges include: defining the optimum biological dose of COX-2 inhibitors in combination with chemotherapy based on clinical and biological, validated surrogate markers.

Discussion – Mark Bernstein, M.D.

Comments: The best administration was when the drug was given every couple of days. More an issue of vincristine not being given often enough to show the improvement.

There is a lot of work suggesting that stroma within tumors are secreting a variety of factors that help with regulation. Therefore, these agents may be affecting the epithelial cells but not affecting the stromal cells that are also helping the proliferation of the tumor.

Q to Dr. Kieran: Is there a chance that the tumor is being depleted of cells that it needs to keep going?

A: There is no question that when any cell type is inhibited, a multitude of effects can be had. There is no single mechanism for all things.

Q to Dr. Healy: Should bisphosphonates be good in lung metastases because it worked in breast, even though those breast patients already had bone marrow metastases; would it be worrisome that it was going after the bone?

A: Yes, but it still reduced both types of metastases. In the orthopedic world these agents are going to be used, but the animal models must be reviewed to see what the data show.

New Target Identification in Osteosarcoma and Model Systems for Testing

cDNA Expression Arrays & Oligonucleotide Expression Arrays - Paul S. Meltzer, M.D., Ph.D.

Dr. Meltzer discussed gene expression profiling for osteosarcoma with cDNA microarrays. With a growing catalog of over 30,000 human genes there is a need to recognize this resource as a catalog of candidate drug targets.

Thus far, from the expression profiling of cancers, it has been learned that distinct cancers have distinct patterns of gene expression. Using expression data, it is possible to develop robust formal diagnostic classifications. Novel subgroups can be recognized which lack previously defined clinico-pathologic correlates. Also, clinical application of array data does not necessarily require arrays; small groups of genes may carry most of the useful information.

Future directions include improving the diagnostic categorization of tumors, identifying useful predictive markers for outcome and therapeutic response (array or conventional), and identifying points for intervention (critical pathways, novel drug targets).

To use microarrays appropriately in a clinical study, proper steps need to be taken when designing trials. The question under investigation should be defined with an appropriate patient sample. There is a need to develop an appropriate and rigorous statistical analysis of array data in order to develop a formal classifier. The design should strive for a result using genes which carry information relevant to the question posed, and allow for validation of the data.

In a pilot study in osteosarcoma, one of the initial goals was to demonstrate whether high-quality microarray data could be generated from osteosarcoma. There were 14 primary osteosarcomas, four Ewing’s Sarcomas, eight osteosarcoma pulmonary metastases, one Ewing’s Sarcoma metastases, and two normal lung samples. There was a near 100% success rate of obtaining evaluable data from this sample set.

The data analysis for bone tumors in the pilot study, after being filtered for quality, showed on cluster analysis a strong cluster of Ewing’s Sarcoma -specific genes and a larger group of osteosarcoma-specific genes. By hierarchical clustering of samples there was significant variability seen with osteosarcoma samples, but not as much variability was seen with normal lung. When viewing the weighted gene list, genes most significantly associated with disease were looked for (these are most highly specific) and ranked by a weighted discriminator method, and a random permutation test was performed.

The results showed there is a very strong difference in signal between osteosarcoma and Ewing’s Sarcoma. There was also a very strong signal difference between osteosarcoma metastases and normal lung. The osteosarcoma metastases are very similar to what is seen in a primary osteosarcoma. Therefore, viewing osteosarcoma metastases vs. osteosarcoma primary may be a good way to see what is upregulated or downregulated during this process. The pilot study sample is not big enough to give good data, but this area should be looked at further.

Dr. Meltzer concluded that high-quality microarray data can be obtained from bone tumors despite their content of extracellular matrix, and clinical correlations should be supported in large studies.

Q/A Session

Q: Why wasn’t osteosarcoma compared to normal bone?

A: No normal bone was available. Regeneration tissue needs to be looked at since bone is a regenerating tissue. With microarrays, it is always better to look at more data from additional samples and relevant tissues.

Q: Have you looked at any small-cell osteosarcomas?

A: No, there were none available to study.

cDNA Expression Arrays & Oligonucleotide Expression Arrays - Deborah E. Schofield, M.D.

Dr. Schofield discussed the analysis of osteosarcoma using oligonucleotide expression arrays. In a study she presented, samples of frozen tissue from 34 primary, childhood or adolescent osteosarcomas, all treated on CCG or German protocols, were analyzed. Follow-up was 3-10 years. There were 18 non-responders and 16 responders. Sample processing was as follows: tissue section and selection, RNA extraction +/- amplification, cDNA synthesis, in vitro transcription and cRNA labeling, and fragment cRNA mix.

Many investigators have been trying to amplify the tissue since there is a lack of samples. Good linear amplification has been demonstrated so perhaps the needle biopsies may supply enough material for these evaluations.

Affymetrix U95Av2 arrays were used. A series of oligonucleotides were synthesized in situ by photolithography, representing known exonic sequences from 3’ to 5’ of ~12,500 genes of interest. Bioinformatics has been key to deciphering significant information from the data output.

Molecular targeting for cancer therapeutics involves the identification of specific gene targets with known biologic function for which therapeutic interventions may be available. The Genetrix tool was used to look at clusters and specific genes. p53 and telomerase expression levels were examined for their relationship to survival and the data showed this to not be a very good predictor of outcome. Her2 expression levels were also looked at, but not as a predictor of outcomes.

For rhabdomyosarcoma, the most expressed gene was FGFR-4 (fibroblast growth factor receptor 4). This gene showed some prognostic significance depending on where the delineation for outcome is defined. Other genes associated with death include: sodium channel 2 (hBNaC2) mRNA, VIP2 receptor KIAA0192 gene (partial cds), PEG3 mRNA (partial cds), RNA polymerase II, 14.5 Kda subunit, autoantigen mRNA, serine/threonine protein kinase EMK, CD44 gene (cell surface hyaluronate receptor gene), and EPH-related receptor tyrosine kinase.

For osteosarcoma, the top gene with the strongest association for survival was Growth Arrest Specific 1 (GAS-1). Other genes associated with survival include: protein tyrosine phosphatase non-receptor type 1, hematopoietic cell-specific Lyn substrate 1, folate receptor 2 (fetal), 47h11 Homo sapiens cDNA, Ta08106 k1 Homo sapiens cDNA, Syndecan T, paired mesoderm homeobox T, ELAV (embryonic lethal abnormal vision), and constitutive endothelial nitric oxide synthase gene. The top gene associated with death was Bone Morphogenetic Protein 4 (BMP-4). Other genes associated with death include: guanine nucleotide binding protein (G protein), sperm associated antigen B, KlAAO751 gene product, Galanin, G protein-coupled receptor 39, translocase of inner mitochondrial membrane 8 (yeast) homolog A, nitric oxide synthase 2A (inducible, hepatocytes), 32g4 Homo sapiens cDNA, and Wnt 10B. CD-31 didn’t show up on the list of genes and has the opposite effect of what would be predicted. Very high levels of CD-31 are associated with survival but only if the cutoff is very high. Tumor suppressing subtransferable candidate 3 showed the strongest association with histologic response to therapy, and tyrosine kinase 2 showed the strongest association with lack of histologic response to therapy.

It is hoped that gene expression profiling may be used to predict prognosis and identify potential novel therapeutic targets in these tumors.

Q/A Session

Comment: Care should be taken. A microarray does not allow viewing of a single gene. Pathways should be looked at. The technique is not designed to look at single genes, rather to look at the pathway.

Comment: The really important issue that biostatisticians have been looking at is sample size requirements for discrimination. The sample size is specific to the array being produced and the hypothesis being testing.

Comment: Yes, that is true. But it is really critical to look at outcome events and first define homogenous sets.

Q: Research has to look beyond the p53 pathway. With Dr. Meltzer’s algorithm, it is not going to be the p53 pathway; it will be much more complicated in osteosarcoma. Technical question: what are the benefits to two different types of arrays being used?

A: The short answer is cost is a big factor with different techniques. It is difficult but not impossible to make a comparison.

Q: What is difference/benefit between Affymetrix and spotted arrays?

A: Very complicated, but Affymetrix appears to give better data.

Comment: Go back to the previous question. A separate set of data for validation is needed. As criteria are developed for how many samples to look at, they should be doubled so they can be validated. Sample size should essentially be made 2x larger so enough samples will be collected to properly to test and validate.

Q: Most of our patients have pre- and post-therapy samples. Looking at histologic response is pretty much subjective, but has what is left as predictive been looked at? Do you have pre- and post- samples?

A: Only a handful, but also have primary and metastases on a handful of patients. Much of what is frozen is necrotic tissue, and it is not very viable tissue.

Q: Is the data ready to give clues as to what should be researched in osteosarcoma?

A: Probably not yet. This is pilot testing data, and it needs to be done on larger groups. Participants will await new data with interest.

Comment: Not every gene is on the chips or is not measured well. So just because the gene isn’t in the results, it may not be on the chip.

Proteomics/CGH - Ching C. Lau, M.D., Ph.D.

Dr. Lau discussed proteomics and CGH.

There are three levels of comprehensive screening targets: genome, transcriptome, and the proteome. The genome is huge and only a small portion actually codes for proteins. The transcriptome (mRNA) is a smaller subset to look at but still problematic because only a subset is expressed in any one cell type at any one time. However, high-throughput technology is available to screen the transcriptome. Finally, purists would argue to view only the proteome because this is a smaller subset and contains only proteins that determine the biological behavior of a cell. Furthermore, there is ample evidence that mRNA expression level may not necessarily correlate with protein expression level or activity due to post-translational modifications and sub-cellular localization. There are many examples where expression level of mRNA is constant but the protein level is variable. Therefore, protein screening theoretically gives you more relevant data. However, proteomics is difficult to carry out, because high-throughput technology is not readily available and existing technology is not very good for quantitation. It is very labor-intensive with low output.

Therefore, RNA screening is currently the most popular and feasible. Dr. Lau briefly discussed their efforts in using cDNA microarrays to generate expression profiles for molecular classification of osteosarcoma. The objectives included: discovering classes of gene expression profiles in samples of primary osteosarcoma; comparing gene expression of primary osteosarcoma to expression of normal osteoblasts; comparing gene expression of osteosarcoma before and after chemotherapy; comparing gene expression of primary osteosarcoma to expression of lung metastases for the same patient, for those patients who present with lung metastases, and in patients who later develop lung metastases; predict response to neoadjuvant therapy; and predict long-term outcome. One of the technical hurdles in analyzing osteosarcoma samples is the small quantity of biopsy specimens and low number of viable cells in the surgical specimens of good responders to neo-adjuvant chemotherapy. Thus RNA amplification is required to minimize sample size requirements but it has to be done carefully in order not to introduce bias into the profile. The T7 amplification schema is favorable for this. Summary of T7 studies: optimized T7 procedure yields >500 fold amplification; 10-20 % are false positives, but very few false negatives (more worried about false negatives so this is acceptable); most of the DE genes from amplified RNA have higher DE ratios than from unamplified RNA (perhaps some of the false positives are actually true positives that become detectable because of increased sensitivity after T7 amplification and can’t be picked up in unamplified studies); and over 90% of the top 100 significant genes were also present in unamplified RNA. Clustering results showed minimal difference between amplified and unamplified samples.

mRNA is not the only way to get data. The CGH method detects gains/losses/amplifications in a single hybridization and maps to a chromosomal region. No cell culture is needed, and archival materials can be used. Limitations include: its limited resolution, its dependence on chromosomal morphology, the labor intensity for data generation, and the need for confirmation by FISH. A significant improvement is to use array CGH which replaces normal metaphase chromosomes with a carefully chosen series of BACs as a platform for detecting chromosomal abnormalities. There is good correlation between the chromosomal CGH vs. array CGH, but the array has higher resolution.

CGH can only detect net gain of genetic materials and should be complemented by spectral karyotyping (SKY). SKY is a 24-color FISH technique that can be used to detect subtle translocation. If you couple SKY with CGH, additional data is given and a better picture of what is going on. 8q24, 12p11, and 14q24 are some of the interesting recurrent regions of abnormalities in osteosarcoma worth looking at.

Finally, Dr. Lau summarized some proteomics data generated using the SELDI ProteinChip Arrays. It can be used for capturing protein in a solid phase and create protein profiles that can be identified by MS-TOF.

Genes Related to Metastases - G. Peter Beardsley, M.D., Ph.D.

Dr. Beardsley discussed genes that are related to metastases. The basic premise of his presentation was that metastasis is a major clinical feature of osteosarcoma and is the proximate cause of death. Therefore, the expression of genes associated with the "metastatic phenotype" in tumor tissue is likely to be predictive of outcome in osteosarcoma. A caution on tumor heterogeneity: a small proportion of tumor cells actually undergo metastasis, and the properties of the gross tumor many not reflect the primary tumor’s true metastatic potential.

The focus of his current project is on the c-met proto-oncogene and matrix metalloproteinases. Cellular responses to c-met ligand bindings activate growth cycle (mitogenesis), start membrane motors (motogenesis), and basement membrane lysis (morphogenesis). c-met immunostaining shows that high c-met expression is a strong negative prognostic indicator in breast cancer (Rimm, et al). Dr. Beardsley and his colleagues at Yale are examining c-met expression levels in osteosarcoma samples using tissue microarray methodology. This technique requires very small amounts of paraffin embedded material, which overcomes a significant problem in tissue acquisition. They are also working on c-met as a therapeutic target).

Matrix metalloproteinases exhibit the following family characteristics: high sequence homology, are enodpeptidases, are Ca++ and/or Zn++ dependent, are involved in the lysis/remodeling of the extracellular matrix, and number more than a dozen. MMP expression shows varied correlations with tumor growth and invasion (good: MMP-2, MMP-9, MMP-14; poor: Matrilysin, stromelysin-3). MMPs as therapeutic targets: Several inhibitors of MMPs are in clinical trials. Although overall results have been disappointing, some activity has been seen in established tumors. They may work best early in therapy to decrease metastasis, and this kind of activity would not be detected in standard Phase I or II clinical trials.

Canine Models - Chand Khanna, D.V.M., Ph.D./Stephen J. Withrow, D.V.M.

Dr. Stephen Withrow presented information on the canine model for bone cancer research. The etiology of canine bone cancer is largely unknown. Large body size and the purity of breed may be contributing factors.

In dogs, more males are affected by osteosarcoma, and 75% of the time it affects the long bones (front leg – 60%, rear leg – 40%). Stage IIB is the most common presentation, and low-grade histology is rarely seen. At presentation, 15% are Stage III. Pulmonary metastases are very common (leading cause of death). Local control is done through NSAIDs, palliative radiation, full-course radiation, amputation, and limb sparing surgery.

There is scientific support of the dog as a model. Osteosarcoma develops spontaneously in dogs, and dogs are large animals that have similar physiology to humans. There is also a rapid case accrual, and autopsy compliance is common. The metastatic rate of osteosarcoma in dogs is similar to humans (80-90%, lung>bone), but time to local and systemic recurrence is quicker in dogs.

Eligibility for treatment on protocol at Colorado State University calls for "localized disease," <50% of bone length involved, a projected one-year survival, and necropsy. Surgery includes "marginal-marginal" excision, allograft replacement, or plate arthrodesis. A treatment regimen model involves cisplatin at Day 1 and 21 , radiation (10fxns, Monday, Wednesday, and Friday), surgery at Day 42, and monthly follow-up. Percent of tumor necrosis and local recurrence mimics the human presentation of the disease.

Prognostic variables that appear to be indifferent for survival are breed, sex, site (except mandible), whether or not limb is spared vs. amputated, and its histologic variant. Poor prognostic variables include young age, metastases at diagnosis, large volume, and high bone ALP. New studies underway include use of chemotherapy (gemcitabine), antiangiogenesis agents, gene therapy, immunotherapy, or bisphosphonates.

The dog model shows potential in causation and epidemiology. Trials can be done in institutions, but need specific resources to make them successful. The limitations of the dog model include: human proteins and immune response to these reagents and therapy; poor canine genome definition; drug/trial costs (good and bad); and toxicity and owner compliance.

Dr. Chand Khanna continued with the biology of canine osteosarcoma. p53 is altered in approximately 50% of tumors. Canine osteosarcoma has RB dysregulations, cytogenetic aberrations, and cDNA microarray cross-reactivity. There will eventually be a dog cDNA microarray, and there is continuing work on the Dog Genome Project. IGF-I inhibition plus chemotherapy in osteosarcoma was investigated. OncoLar was used to see if it could be moved to pediatric patients in combination with chemotherapy. "In vivo canine with OncoLar and carboplatin" used OncoLar + chemo + amputation in one arm, and chemo + amputation in the other arm (a great source of tissue is from amputation). Among 64 cases in 8 months, with survival as the clinical endpoint, there was no difference in necrosis as a primary tumor response, no significant difference in apoptosis between OncoLar and placebo group, and survival showed no improvement in dogs receiving OncoLar. This combination may not represent the best treatment approach with OncoLar.

Ongoing and future projects include: examining the influence of 70% serum IGF-I suppression on osteosarcoma primary tumor growth and metastasis; examining local regulation and production of IGF-I and GH in osteosarcoma; and evaluation of GH antagonists and GH-RH antagonists (murine K-series osteosarcoma model and canine osteosarcoma).

The dog model is quite valuable and hopefully can become part of the standard preclinical approach to therapy development.

Metastases Associated Genes - Chand Khanna, D.V.M., Ph.D.

Dr. Khanna presented additional information on the genetic determinants of osteosarcoma metastases.

In the murine osteosarcoma model, K7M2 has a more aggressive and K12 has a less aggressive metastatic potential. The murine microarray analysis, a 4k array, looked at 59 differentially expressed genes from K7M2 primary tumor (34 genes overexpressed) and K12 primary tumor (25 genes overexpressed). K7M2 shows greater genetic expression of genes related to cytoskeletal motility, adherence, and angiogenesis than K12.

Ezrin protein is expressed at the cell membrane in the more aggressive cells. These proteins are phosphorylated at the membrane where they are activated, and are important for signaling and interaction. A tissue array using Ezrin in dogs with osteosarcoma showed that metastases have a statistically higher level of Ezrin expression compared to primary tumor. Dogs with any level of Ezrin expression have poorer survival than dogs not expressing the protein, but this survival difference is not statistically significant. As for Ezrin in human osteosarcoma, there also may be an association with metastases. Exploratory results were somewhat disappointing, as Ezrin expression was detected in human osteosarcoma, but expression was not clinically distinctive.

For ongoing and future work, the murine osteosarcoma model will be utilized for the examination of the biology of Ezrin mutants in vitro/in vivo and Ezrin regulation during metastasis. For the spontaneous canine osteosarcoma, development of larger tissue arrays and reevaluation of phosphorylated Ezrin expression will be explored. For pediatric osteosarcoma, increasing sample size (metastases at diagnosis), reevaluating phosphorylated Ezrin expression, and standardizing tissue collection will be investigated.

Evaluation of Camptothecins, Epothilones and Signaling Inhibitors in Osteosarcoma Models - Peter Houghton, Ph.D.

Dr. Houghton began his talk by stating he believed that the best types of models would be useful for chemotoxic as well as molecularly targeted agents. Animal models of osteosarcoma can be spontaneous, transplanted (syngeneic), transplanted (xenogeneic), or orthotopic. A perfect model does not exist, but one needs to be found that parallels sensitivity to human cancer, parallels the genetic basis for human cancer, parallels the biology of human cancer, and allows for high throughput for drug screening.

Xenograft models of childhood cancer include rhabdomyosarcoma, neuroblastoma, Ewing’s Sarcoma, brain tumors, Wilms’ Tumor, and osteosarcoma. Animal models of cancer can help in understanding the genetic basis of cancer, developing cancer prevention and cancer treatment techniques, identifying carcinogens, and in drug development/PK/safety.

Preclinical development/validation issues with osteosarcoma models include the following: does the model identify clinically active agents prospectively; does the model encompass known clinical heterogeneity; what is the basis for false prediction; and can we develop approaches that minimize false predictors.

Clinically, there seems to be relative agreement with xenografts. Irinotecan (CPT-11) is somewhat active in osteosarcoma xenografts, but these xenografts are much less sensitive than rhabdomyosarcoma, neuroblastoma, or Wilms’ Tumor models. Signaling inhibitors (Gleevec, Iressa) do not have significant single agent activity; and Epothilone B (BMS247550) has some activity. Dr. Houghton concluded that osteosarcoma xenografts may be useful models for identifying new agents and combinations. They are generally less sensitive than other childhood tumors to camptothecins. Furthermore, orthotopic models may have some value, specifically to lung metastases.

Guest Presentation: Activation of a Redundant Caspase Pathway in Bones of p53 Null Mice - Janet M. Hock, B.D.S., Ph.D.

Osteosarcoma is thought to be a tumor of osteoprogenitor cells: multipotential, hormone-responsive stromal cells in the periosteum and marrow, which are capable of differentiating into many lineages depending on their environmental cytokines. As in other tumors, the p53 gene has been linked to increased susceptibility to osteosarcoma. Published studies suggest that p53 may have a role in the normal development and physiology of bone. Failures in skull growth and delayed longitudinal bone growth have been described in utero in p53 null mice. We undertook an extensive characterization of bone, including hormonal responsive to parathyroid hormone, in p53 null mice aged 6 and 12 weeks. Once daily parathyroid hormone, which has anabolic effects on bone mass and strength, promoted osteosarcoma in rats after 18 months of treatment in a 2-year toxicology study. Although the relevance of PTH-induced rat osteosarcoma to humans is not known, it may provide an animal model for studies of bone sarcomagenesis.

Our data showed that p53 null mice had smaller, thinner bones than their wild-type counterparts. The smaller bones were associated with decreased surface bone formation. Premature closure of sutures at the base of the spine was associated with loss of symmetry in the maxilla, suggesting a key role for p53 in normal development of the craniofacial complex. When challenged with PTH, expression of the AP-1 complex of genes, specifically c-fos and fra-2, was blunted in p53 null mice while in vitro induction of osteoclast differentiation was enhanced. As p53 is a key regulatory gene determining apoptosis and PTH inhibits apoptosis, we investigated bone cell caspase activity as a surrogate marker of apoptosis. PTH inhibited caspase activity in both p53 intact and null mice. However, p53 null mice had switched to an alternate caspase pathway as activity was seen on YVAD substrate, which recognizes caspase 1 (interleukin-converting enzyme-beta) rather than on DEVD substrate that recognizes the apoptotic caspases 2, 3 and 7. The implication of this change in caspase pathways on skeletal function and osteosarcoma susceptibility is unknown.

Q: How hard is it to get osteoblasts?

A: It depends. Early progenitors are in bone marrow, and they also can come out of other bone that is more advanced, but those aren’t as good in culture.

Discussion - Paul A. Meyers, M.D.

Q to Dr. Grier: The dog model is very important. The question is how to, or what organization, can gather more dogs. Is there further structure that is needed or more money? Is the money there from NCI?

A: The dog is a good model, but funding from an R01 mechanism is very difficult. People don’t see the relevance; osteosarcoma is a rare disease, and the research is in dogs.

A by Dr. Khanna: The way to move forward is to make the dog model part of a larger program. Defining questions that are important in this type of group could lead to a P01.

A by Dr. Toretsky: The model is very interesting, although, it is unknown how many of the other drugs correlate with human response. Is it going to teach about biology and replicate what would be seen in humans? Probably a directed RFA would be important.

Comment: Whether the model is predictive or not is dependent on the agent. Some agents are predictive and others may not be. It looks like the dog model is part of a component about how to get better ideas of where to be going. As for money, it would be helpful to describe what research should be done, and so the funds can be properly allocated.

Comment: The canine model seems great. It can be further enhanced, though, to enhance PK studies that go along with the model. Some of our plans should focus on this.

Concluding Remarks - Barry D. Anderson, M.D., Ph.D., and Richard Gorlick, M.D.

Pathogenesis of Osteosarcoma

Lessons:

•   Osteosarcoma likely results from a variety of genetic alterations far more complicated than many other pediatric malignancies
•   No signature translocations have been identified
•   Complex karotypes indicate tremendous genetic instability
•   Subtypes of osteosarcoma may be distinguished by viral or suppressor gene alterations
•   SV40 sequence in some tumors – T-Ag expression
•   Blockade of p53/RB function by T-Ag
•   Osteosarcoma pathobiology may evolve through a series of genetic alterations or cell regulatory modifications
•   A correlation between telomerase and aggressiveness remains unclear

Future Directions:
•   Biology samples to be collected or processed?
Discussion:

Should we look at mismatch repair genes? People expressed concern that we not look at one gene at a time. Massive genomic instability is the common genetic finding in osteosarcoma and this is a biologic issue that has not been explored. There are fairly straightforward ways to screen for mismatch repair. Should we incorporate people who have experience with DNA repair in this future work? This approach was questioned because an association of repair deficit syndrome was not known.

What type of samples need to be collected? Is all paraffin embedded the same? How is it processed? Are you trying to look at which studies you can do with tissue you already have or what new studies can do and how to get more tissue? Both. Tissue samples are becoming available and we need to know what to study with them and how to best collect them. Need to figure out best use of tissue since it is not an unlimited resource.

In terms of paraffin blocks, do you collect this? Maybe paraffin can be used for validation studies rather than primary studies. The COG 9851 protocol requirements include paraffin embedded tissue. The type of paraffin fixation is not specified, as this is viewed as the domain of the pathologist to specify this. The other issue is storage of the slides.

Cytokine/Death Receptor Pathways in Osteosarcoma

Lessons:
•   Influence of Fas/FasL in osteosarcoma pulmonary metastases (with possible explanation of INT-0133)
•   Multiple (and complicated) pathways to tumor control in the IL-12/pulse IL-2 therapy
•   Knowledge gained by looking at multiple cell systems and fresh tumor in Ewing’s Sarcoma and neuroblastoma
•   Ability to move toward cytokine/chemotherapy combination trials in colon cancer

Forward Movement:
•   Further work in osteosarcoma animal models
•   Further characterization of cytokine/cytokine receptors in the SAOS-LM6 model
•   Confirmation in other osteosarcoma lines
•   Fas/FasL in patients at presentation
•   Cytokine changes with aerosolized GMCSF study
•   Clinical background for possible future rx with aerosolized IL-12

Drug Resistance/Growth Factor Pathways in Osteosarcoma

Growth Factor Pathways:
•   Antibody-based strategies for targeted therapy are interesting, but too preliminary for osteosarcoma studies
•   IGF-1/GH – preliminary data suggest it might have a role in tumor etiology
•   P9851 might help better define the role of IGF-1 in osteosarcoma
•   Further preclinical evaluation is important to attempt to define its role as a therapeutic target

Discussion:

A participant added that before things are dismissed, you must understand how to measure success. Survival may incorporate too many things, and you may not exactly see the potential of the agent. Future symposium should have a section on the evaluation of endpoints to properly measure effectiveness of drug. But survival still has to be the gold standard. Others can be looked at, but survival is the most important.

Small molecule inhibitors for IGF-I may be available within 2-5 years.

P-glycoprotein is a low priority.

Comment on prioritization and use of samples. Use NWTSG as an example. They also have a reserve bank that no one touches unless it is high priority. Should also use their ideas to correlate markers with lower numbers.

New Treatment Approaches

New Treatment Approaches:
•   Concepts and Protocols in Development
•   Samarium: for patients with bone recurrence
•   Inhaled GMCSF: for patients with pulmonary only recurrence
•   Trimetrexate: window study for patients with disease metastatic at initial diagnosis, not eligible for Herceptin study
•   Viral approaches: needs further development work - MSK study
•   Slow release platinum: potentially interesting idea for incorporation: efficacy data in dogs, need safety and preliminary data in humans, discussion on priority
•   Antiangiogenic approach: in development for Ewing’s Sarcoma metastatic at diagnosis (vinblastine 1 mg/m2 3X/week with celecoxib 250 mg/m2 BID), as a backbone for standard cytotoxic treatment; squalamine could be incorporated with cisplatin
•   Bisphosphonates: data collection

Slow release platinum: it is currently unavailable; maybe Peter Anderson can look at it in xenograft model to get more pharmacology information on it. Combination with radiotherapy may work to get it into a phase I trial...possibly head and neck. The anti-angiogenic approach is not ready to advance. Bisphosphonates still need data. Someone suggested using a canine model to test all of these as a single agent and then add bisphosphonates.

New Target Identification in Osteosarcoma and Model Systems for Testing

Discussion:

All meeting participants were supporters of array technology. A further look at canine models was encouraged. If Peter Houghton can identify highly valuable compounds, then there is a need to pursue the dog model and increase testing in xenograft models.

MEETING ADJOURNED


Summary

Published in Clin Cancer Res. 2003 Nov 15;9(15):5442-53. Full text article at http://clincancerres.aacrjournals.org/cgi/content/full/9/15/5442 .

Richard Gorlick1, Peter Anderson2, Irene Andrulis3, Carola Arndt2, G. Peter Beardsley4, Mark Bernstein5, Julia Bridge6, Nai-Kong Cheung1, Jeffrey S. Dome7, David Ebb8, Thomas Gardner9, Mark Gebhardt8, Holcombe Grier10, Marc Hansen11, John Healey1, Lee Helman12, Janet Hock9, Janet Houghton7, Peter Houghton7, Andrew Huvos1, Chand Khanna12, Mark Kieran10, Eugenie Kleinerman13, Marc Ladanyi1, Ching Lau14, David Malkin15, Neyssa Marina16, Paul Meltzer1, Paul Meyers1, Deborah Schofield17, Cindy Schwartz18, Malcolm A. Smith22, Jeffrey Toretsky19, Maria Tsokos12, Leonard Wexler1, Jon Wigginton12, Stephen Withrow20, Mason Schoenfeldt21 and Barry Anderson22

1 Memorial Sloan-Kettering Cancer Center, New York, New York;
2 Mayo Clinic, Rochester, Minnesota;
3 Mount Sinai Hospital and Research Institute, Toronto, Ontario, Canada;
4 Yale University School of Medicine, New Haven, Connecticut;
5 Hopital Sainte-Justine, Montreal, Quebec, Canada;
6 University of Nebraska Medical Center, Omaha, Nebraska;
7 St. Jude Children’s Research Hospital, Memphis, Tennessee;
8 Massachusetts General Hospital, Boston, Massachusetts;
9 Indiana University Medical Center, Indianapolis, Indiana;
10 Dana-Farber Cancer Institute and Children’s Hospital, Boston, Massachusetts;
11 Health Sciences Center, University of Connecticut, Farmington, Connecticut;
12 Pediatric Oncology Branch, National Cancer Institute, Bethesda, Maryland;
13 M. D. Anderson Cancer Center, Houston, Texas;
14 Texas Children’s Cancer Center at Baylor College of Medicine, Houston, Texas;
15 Hospital for Sick Children, Toronto, Ontario, Canada;
16 Stanford University Medical Center, Stanford, California;
17 Children’s Hospital Los Angeles, Los Angeles, California;
18 Johns Hopkins Hospital, Baltimore, Maryland;
19 Georgetown University Medical Center, Washington, DC;
20 College of Veterinary Medicine and Biological Sciences, Colorado State University, Ft. Collins, Colorado;
21 EMMES Corporation, Rockville, Maryland; and
22 Cancer Therapy Evaluation Program, National Cancer Institute, Rockville, Maryland

Abstract

Childhood osteogenic sarcoma (OS) is a rare bone cancer occurring primarily in adolescents. The North American pediatric cooperative groups have performed a series of clinical treatment trials in this disease over the past several decades, and biology studies of tumor tissue have been an important study component. A meeting was held in Bethesda, Maryland on November 29–30, 2001, sponsored by the NIH Office of Rare Diseases, the Children’s Oncology Group, and the National Cancer Institute-Cancer Therapy Evaluation Program with the general objectives: (a) to review the current state of knowledge regarding OS biology; (b) to identify, prioritize, and support the development of biology studies of potential clinical relevance in OS; and (c) to discuss the available tissue resources and the appropriate methods for analysis of OS samples for the conduct of biology studies. This report summarizes the information presented and discussed by the meeting participants.

Introduction

OS23 biology studies are becoming increasingly important because of the clinical need for prognostic factors to stratify treatment and for new molecular targets that could indicate a role for targeted therapeutic agents in OS treatment. A meeting was held in Bethesda, Maryland, on November 29–30, 2001, sponsored by the NIH Office of Rare Diseases, Children’s Oncology Group, and the National Cancer Institute-Cancer Therapy Evaluation Program with the following broad general objectives: (a) to review the current state of knowledge and share information regarding OS biology; (b) to identify, prioritize, and support the development of biology studies of potential clinical relevance in OS; and (c) to discuss the available tissue resources and the appropriate methods for analysis of OS samples for the conduct of biology studies. This article will highlight some of the information presented at the meeting.

Pathogenesis of Osteogenic Sarcoma
Genetic Abnormalities.

It is useful to place OS in the context of other sarcomas. In terms of oncogenetic mechanisms, there seem to be two major classes of sarcomas: (a) sarcomas such as ESFTs, with balanced karotypes (reciprocal translocations) and alterations of tumor suppressor gene pathways in relatively few cases; and (b) sarcomas such as OS, with complex unbalanced karotypes, and alterations of the p53 and retinoblastoma pathways in most cases (1). Despite the complexity of the karyotype and the absence of characteristic reciprocal translocations, many recurrent, nonrandom chromosomal abnormalities are observed in OS.

A University of Nebraska Medical Center study analyzed 128 OS specimens, of which 70% were clonally abnormal (2, 3). Most demonstrated pronounced heterogeneity within the same patient. Marker chromosomes, structurally abnormal chromosomes in which no part can be identified, were detected in the majority of OS samples (58%). Ring chromosomes (7%) accompanied by multiple numerical (65%) and structural (72%) abnormalities were also prominent (2, 3, 4). There is evidence of genomic amplification (homogeneously staining regions or double min) in at least one-third of the cases. Cytogenetic abnormalities observed were both numerical and structural. Common numerical abnormalities in OS include: gain of chromosome 1; loss of chromosomes 9, 10, 13, and/or 17, and partial or complete loss of the long arm of chromosome 6 (2, 3, 5).24 Frequent structural abnormalities include rearrangements of chromosomes 11, 19, and 20 (2, 3, 5, 6).24

Tumor Suppressor Pathway Alterations.
The p53 and retinoblastoma tumor suppressor pathways are clearly involved in the pathogenesis of OS. Most OS samples have some type of combined inactivation of the retinoblastoma and p53 tumor suppressor pathways (7). In a study of 32 OS specimens, a number of loci were demonstrated to have LOH (3q, 13q, 17p, and 18q), including the locations of the Rb and p53 tumor suppressor genes (8, 9, 10). The timing and sequence of these alterations, particularly relative to the development of the chromosomal complexity, and other tumor suppressor and oncogene alterations, are unclear. Evidence for the role of p53 in OS pathogenesis includes the predisposition of patients with germ-line p53 mutations to develop OS (11, 12, 13).

Amplification of the 12q13 region (containing MDM2 and CDK4) or INK4A deletion can affect both the p53 and Rb pathways, and, indeed, these alterations seldom coexist with Rb or p53 alterations (14). Because these tumors almost universally have genetic alterations that inactivate the Rb and p53 tumor suppressor pathways, gene inactivation by itself may not be a strong prognostic factor. Indeed, new data show low prognostic significance of p53 mutations in sporadic OS. In a recent study (15), 22% of OS samples showed p53 mutations, but there was no relationship to distant recurrence. Interestingly, p53 status was concordant in all of the paired samples of primary and distant metastases, suggesting p53 pathway alterations may occur early in OS pathogenesis.

Genome-wide attempts have been made to identify potential tumor suppressor genes associated with the LOH in OS (16, 17). Examination of 38 chromosomal arms from OS tumor samples for LOH has found that the mean frequency of LOH is 30.79% for any chromosome arm, an unusually high mean frequency for a childhood tumor. Moreover, several chromosome arms (3q, 13q, 17p, and 18q) underwent LOH with a frequency >2 SDs higher than the average (P < 0.002; Ref. 16). Further mitotic mapping has identified minimal regions thought to contain candidate tumor suppressor loci on chromosomal arms 3q26.2 and 18q21.33 (9, 10). Additional analysis has suggested that other chromosomal regions may also harbor tumor suppressor loci important in OS tumorigenesis including chromosome arms 5q, 6q, 10q, 11p15q, 16p, and 22q (18).

OS is thought to be a tumor of osteoprogenitor cells. These are multipotential, hormone-responsive stromal cells in the periosteum and marrow, which are capable of differentiating into many lineages depending on their environmental cytokines (19). Published studies suggest that p53 may have a role in the normal development and physiology of bone (20). Failures in skull growth and delayed longitudinal bone growth have been described in utero in p53 null mice. An extensive characterization of bone, including hormonal responsiveness to PTH, was performed in p53 null mice (21). The p53 null mice had smaller, thinner bones than their wild-type counterparts; these changes were associated with decreased surface bone formation. Premature closure of sutures at the base of the skull was associated with loss of symmetry in the maxilla, suggesting a key role for p53 in normal development of the craniofacial complex. Daily PTH administration, which has anabolic effects on bone mass and strength, promoted OS formation in rats after 18 months of treatment in a 2-year oncogenicity study (21). When challenged with PTH for a few days, expression of the AP-1 complex of genes, specifically c-fos and fra-2, was blunted in p53-null mice, whereas in vitro induction of osteoclast differentiation was enhanced (22). Because p53 is a key regulatory gene determining apoptosis, apoptosis was investigated in bone cells isolated from long bones of p53 intact and p53 null mice, using substrates degraded by caspases as surrogate markers for apoptosis. Confirming previous data (23), bone cell lysates from intact mice degraded DEVD substrate, which recognizes the apoptotic caspases 2, 3, and 7, and showed no significant activity with substrates recognizing caspases 1, 4, 5, 6, 8, or 9. In contrast, bone cell protein lysates from p53-null mice exhibited substantial activity on caspase substrates recognizing caspase 1 activity and only minimal activity on DEVD substrate. This suggests that in the absence of p53, bone cells switch to an alternate pathway that activates caspase 1 (interleukin-converting enzyme-ß) rather than caspases 2, 3, and 7. PTH inhibited the caspase 3 activity in intact mice and the caspase 1 activity in p53 null mice. Although the p53-dependent switch from apoptotic caspase 3 activity to caspase 1 activity could indicate existence of a redundant pathway, caspase 1 is usually considered nonapoptotic, because it cleaves cytokines associated with inflammation and osteoclast induction. The implication of this change in caspase pathway on OS induction is unknown.

Telomerase and ALT.
The complex unbalanced karyotypes that characterize OS may reflect its pathogenesis. Karyotypic complexity may reflect chromosomal fusion-bridge-breakage cycles that occur due to advanced telomere erosion. A potential etiology of this chromosomal instability is telomere dysfunction, as has been implicated in epithelial cancers (24). Telomeres are nucleoprotein structures that cap chromosome ends and serve at least three protective functions: (a) preventing chromosomes from being recognized as damaged DNA; (b) preventing chromosomal end-to-end fusions and recombinations; and (c) accommodating the loss of DNA that occurs with each round of replication. Normal human somatic cells have finite proliferative capacity, and it has been demonstrated that telomere length is one of the checkpoints that determines when a cell stops dividing (25, 26). As cells divide, the telomere length gradually decreases to a critical size, at which point senescence is triggered by a p53-dependent process. Human cells may bypass this checkpoint by inactivating the p53 pathways, in which case they continue to divide until telomeres become very short, chromosomal instability ensues, and apoptosis is triggered. Rare cells bypass this second checkpoint by activating mechanisms that lengthen telomeres, a central feature of cancer cells. About 85% of cancers activate an enzyme called telomerase, which lengthens telomeres, and the other 15% of cancers use a recombination-based method called ALT (27, 28). Unlike most cancers, at least 50% of OS samples are dependent on the ALT mechanism to maintain telomeres (29, 30, 31). The ALT and telomerase mechanisms are different means to the same end, but they are not equivalent. Telomerase-dependent OS cell lines have short telomeres with a minor range of length, whereas ALT-dependent OS cell lines have long telomeres with great heterogeneity in length. The ALT cell lines also have greater genetic instability and more translocations than the telomerase-positive cell lines (29). In a mouse model, five of five ALT-dependent cell lines were unable to generate macroscopic lung metastases, despite robust s.c. tumor growth (32). Upon telomerase reconstitution, all five of the cell lines formed massive pulmonary nodules after tail-vein injection, indicating diminished metastatic potential in ALT-dependent versus telomerase-dependent tumors. It has been hypothesized that ALT-dependent human OS have different clinical behavior than telomerase-dependent OS, but this remains to be studied.

Viral Pathogenesis.
SV40, a DNA polyomavirus, consists of three major nonstructural proteins: (a) the small T antigen (enhances the ability of large T antigen to transform cells); (b) the smaller T antigen (function unknown); and (c) the large T antigen (assists in viral replication, interacts with p53 and Rb, and promotes cell proliferation by inhibiting p53 function; Ref. 33). A 1996 study concluded that 11 of 18 OS samples showed evidence of incorporated SV40 DNA (33); a 1998 study showed no correlation of presence of SV40 with p53 or Rb mutation status (34); and a 1997 study showed 50% of OS samples tested had incorporated SV40 DNA from each of the four regions of the viral genome (35). Although SV40 is not observed in all tumors, it may exist in select families affected by OS. It is postulated that SV40 large T antigen may inactivate the wild-type p53 allele in the osteoblasts of susceptible individuals, resulting in tumor development. In two families with Li-Fraumeni syndrome, in which family members with cancer harbored germ-line heterozygous p53 mutations, SV40 large T antigen was amplified and expressed in tumor cells that retained the wild-type p53 allele (36).

Future needs and directions in the characterization of the molecular pathology of OS include: (a) incorporating the current lists of genetic alterations into functionally related groups of genetic alterations (hyperproliferative, cell cycle control, apoptosis, and DNA damage response); (b) a better understanding of the timing and relationship of common oncogenetic events in OS [alterations of p53 and Rb as early preclinical events, p53 pathway alterations in OS in familial retinoblastoma (37), and retinoblastoma pathway alterations in OS in Li-Fraumeni syndrome]; (c) developing a comprehensive analysis of the p53 and retinoblastoma pathways in a single large set of OS samples by new high throughput approaches; (d) a better understanding of different "equivalent" common oncogenic events (preferential 12q13 amplification in low-grade/surface OS and preferential p53 missense mutation in adult OS); (e) a better understanding of the paradox of carcinoma-type cytogenetics in the setting of a younger age range; and (f) defining the biological/genetic subsets of OS according to telomerase status and karyotypic complexity.

Cell Death/Cytokine Pathways.
The Fas cell death pathway has been implicated as having a role in determining chemosensitivity and metastatic behavior in a variety of tumors including OS, ESFT, and neuroblastoma (38, 39). Fas is a type I transmembrane protein that regulates immune cell apoptosis, but Fas is expressed in many different cells and tissues, and is functional in cells outside of the immune system (40). Tumor cells expressing surface Fas will apoptose when FasL is presented unless a mechanism of resistance is present. Mechanisms of resistance include down-regulation of Fas, FasL, or procaspase 8 (FLICE), and up-regulation of apoptosis inhibitory proteins, specifically bcl-2 or FLICE-inhibitory protein. Some tumors express FasL on their cell surface enabling a "counterattack" against activated T cells to escape immune surveillance (41). FasL has been shown to exist in transmembrane and soluble forms in ESFT. It has been shown in certain circumstances that ESFT cells may avoid apoptosis by down-regulating their surface FasL (by cleavage) and Fas (by internalization); however, treatment of ESFT cells with synthetic matrix metalloproteinases increased surface FasL and Fas level (in the absence of increased transcription) expression, and decreased cytoplasmic Fas (42).

Fas signaling may be involved in OS metastases (43). A cell line, SAOS-LM6, has been developed through repetitive outgrowth of the pulmonary metastases that arise after SAOS-2 tail-vein injection (44), and SAOS-LM6 has a greater tendency to form pulmonary metastatic lesions. Fas expression was examined in SAOS-LM6 to determine its role in OS metastases. Using Northern analysis, it was found that SAOS-LM6 expressed a lower level of Fas as compared with the parental SAOS-2 line, and flow cytometry demonstrated that this decreased expression was reflected in the cell surface protein (45, 46, 47). To test whether Fas affects the metastatic potential of the cell lines, Fas-transfected SAOS-LM6 cells and SAOS-2 control cells were injected into mice. The mice injected with control cells developed metastases, but the mice with Fas-transfected cells did not. This demonstrates Fas expression may be important to the metastatic potential of OS. Additional studies were undertaken to determine whether IL-12, which can up-regulate Fas expression, can also alter the metastatic potential of OS cell lines. IL-12 was transfected into SAOS-LM6 cells using coincubation with an adenoviral vector, and this reduced the metastatic potential of the transfected cells, again supporting the involvement of Fas in OS metastases. Murine IL-12 can be induced in lung tissue by intranasal administration of Ad.mIL-12 vector. A study was conducted in which mice were injected with tumor cells, and then aerosolized Ad.mIL-12 was administered. The mice receiving the Ad.mIL-12 developed no (4 of 8 mice) or <12 (4 of 8 mice) pulmonary nodules, whereas mice treated with a control adenoviral vector developed 25 to >200 nodules in 5 of 8 mice (45, 46, 47).

The Fas pathway may explain the clinical results of a recent pediatric cooperative group clinical Phase III trial (48) in nonmetastatic OS in which patients received doxorubicin, cisplatin, and methotrexate and were randomized to treatment with ifosfamide and/or MTP-PE, or doxorubicin, cisplatin, and methotrexate alone. MTP-PE can induce IL-12 in patients through activation of macrophages. Ifosfamide may up-regulate FasL expression on OS tumor cells in a manner analogous to the in vitro effect of cyclophosphamide in the SAOS-LM6 and SAOS-LM6-IL-12-transfected OS cells. The combined induction of both IL-12 and FasL could result in enhanced tumor apoptosis, and this Fas-related event may explain the trend toward improved outcome observed in patients treated with both ifosfamide and MTP-PE compared with either individual agent.

Clinical investigations of IL-12 combined with IL-2 have been initiated in adult cancers (49). The rationale for expecting complementary antitumor effects by IL-12 and IL-2 is as follows: different signaling pathways are used by IL-12 versus IL-2; and IL-2 up-regulates IL-12 receptor expression on T and natural killer cells, and the combination has enhanced immune-stimulating effects for T and/or natural killer cells (cytotoxicity, cytokine production, and proliferation). A regimen of systemic IL-12 and pulse IL-2 was investigated using murine models of renal cell (50) and mammary (51) carcinoma. The regimen was well-tolerated and resulted in complete tumor regression in 90–100% of mice with metastatic renal cell carcinoma and 50% of mice with mammary carcinoma. Systemic IL-12/pulse IL-2 inhibited tumor neovascularization and induced tumor regression via mechanisms that were dependent on CD8+ T cells, IFN-γ, and Fas/FasL (52). Additionally, investigators observed that most mice cured of the renal cancer were resistant to rechallenge with the tumor, and the tumor became infiltrated with CD8+ T cells. The regimen was ineffective in SCID mice, and depletion of the CD8+ T cell effector cells ablated the antitumor response, suggesting the importance of an immune mechanism. The impact of IL-12/pulse IL-2 on tumor vascularization may occur via IFN-γ-dependent induction of Fas/FasL, and ultimately, vascular endothelial cell apoptosis. With this preclinical data, Phase I studies of systemic IL-12/pulse IL-2 have been initiated in adults with advanced solid tumors. Along with these findings, the subsequent demonstration of potent therapeutic efficacy by IL-12 ± IL-2 in orthotopic models of murine neuroblastoma has provided preclinical rationale for a Phase I investigation of IL-12 ± pulse IL-2 that has now been in initiated in children with persistent or refractory neuroblastoma.

Drug Resistance Pathways.
OS tumor recurrences are likely to be at least partly due to drug resistance (53). Possible mechanisms of drug resistance include alterations in: p-glycoprotein expression, multidrug resistance protein expression, topoisomerase II, glutathione S-transferases, DNA repair, drug metabolism or inactivation, and reduced intracellular influx. Drug resistance can be intrinsic (in the absence of treatment) or acquired (after treatment with chemotherapeutic agents; Ref. 54).

P-Glycoprotein Expression.
Of the drug resistance mechanisms, p-glycoprotein expression has been the most extensively studied in OS. P-glycoprotein is an ATP-dependent efflux pump for hydrophobic substances, and its expression can result in resistance to doxorubicin and etoposide. P-glycoprotein expression in OS has been investigated as a means to identify "nonresponders" early, allow treatment modification, and potentially improve their survival. Similarly, negative p-glycoprotein expression may identify patients who are more responsive and could receive less toxic therapy.

Published studies have concluded that p-glycoprotein levels can predict poor outcome, defined as development of metastases and death, in OS (55), associating p-glycoprotein expression with an adverse outcome risk ratio of 3.37 (95% confidence interval, 1.6–7.1). A review of the many published studies of p-glycoprotein in OS (55, 56, 57, 58, 59, 60, 61, 62, 63, 64) suggests that p-glycoprotein expression is associated with a poorer metastasis-free survival and poorer overall survival. However, the immunohistochemical measurement of p-glycoprotein expression in the tumor biopsy samples of patients participating in the pediatric OS Phase III study discussed previously (65), did not detect a significant difference in event-free survival (relative risk = 1.05) or risk of death (relative risk = 1.0) between patients who demonstrated antibody positivity at study entry compared with antibody-negative patients. This study concluded that p-glycoprotein expression is not useful in determining the prognosis of patients presenting with localized OS treated with cisplatin, methotrexate, doxorubicin, ± MTP-PE, ± ifosfamide. Furthermore, tumor histological response to neoadjuvant chemotherapy did not correlate with p-glycoprotein positivity.

Caveats of studies of multidrug resistance in OS include: (a) the decreased survival associated with p-glycoprotein-positive tumors may not reflect decreased drug accumulation; (b) the presence of p-glycoprotein does not necessarily confer resistance; and (c) its absence does not imply sensitivity. In addition, it is known that other efflux proteins such as multidrug resistance protein are present in OS. The current literature concludes that there is no correlation between p-glycoprotein status and percentage of OS tumor necrosis after induction chemotherapy. The literature also suggests that p-glycoprotein may be a sign of aggressiveness as well as a marker of drug resistance, and may be useful in identifying high-risk OS patient subsets (66).

MDR1 Expression.
In a pilot study of 15 patients, the expression of MDR1, which encodes the protein p-glycoprotein, was correlated with poorer overall survival. In a larger prospective study, MDR1 expression was examined using reverse transcription-PCR and then correlated with disease-free survival. Between 1989 and 1994, 123 newly diagnosed patients with high-grade, nonmetastatic extremity OS all received neoadjuvant chemotherapy, including doxorubicin. Although tumor size was predictive of outcome, the degree of tumor necrosis was not significant in the multivariate analysis. Patients with high levels of MDR1 did poorly, but patients with low levels of MDR1 also did poorly. Therefore, no association between survival and MDR1 could be made (relative risk ratio 1.15; Ref. 63). Microarray analyses or continuous monitoring of MDR1 levels may allow a determination of the reasons patients with both very low and very high levels of MDR1 expression have a poor outcome. Other reasons for the lack of correlation between MDR1 expression levels and disease-free survival in this study as compared with previous may include: conflicting results with immunohistochemistry due to sensitivity of technique, antibodies detecting other proteins, expression of MDR2 in OS, other mechanisms of drug resistance, and the relationship between p53 and drug resistance. Many OS specimens have very high levels of expression of both MDR1 and MDR2. Wild-type p53 can inhibit MDR1 expression (67). When progressive sets of tumor biopsies from the same patient were examined at diagnosis and definitive surgery, it was seen that if the biopsy had a p53 mutation, then the resection had a mutation. When correlating type of p53 mutation and MDR1 expression, it was found that those who had missense mutations had higher MDR1 expression, and those who had nonmissense p53 mutations had lower levels of MDR1.

Antifolate Resistance.
Studies of antifolate resistance have been performed in OS (68, 69, 70). Most soft tissue sarcomas are intrinsically resistant to methotrexate secondary to impaired polyglutamylation (68). The literature suggests that the peak methotrexate level, and not the systemic exposure, correlates with therapeutic response in OS (71, 72). The peak serum level is predominantly determined by drug dose. This suggests that impairments in drug influx may be a basis of methotrexate resistance in OS. Studies of methotrexate resistance have been performed on biopsy samples to identify intrinsic mechanisms of resistance and on relapsed samples to identify acquired mechanisms. Studies have shown that impairments of drug influx secondary to decreased expression and mutations in the reduced folate carrier gene, the major membrane transporter of methotrexate into cells, is the major basis of intrinsic resistance (69). Dihydrofolate reductase, the target of methotrexate, overexpression is the major mechanism of acquired resistance (69).

Trimetrexate, a newer antifolate, has been investigated as an alternative approach for OS treatment based on its ability to enter into cells that lack a functional reduced folate carrier (73). A Phase II trial of simultaneous trimetrexate and leucovorin given orally for 21 days was conducted in relapsed OS patients. Toxicity was acceptable with myelosuppression being the major side effect, and objective responses were observed in 8% (n = 39, complete response = 1, partial response = 2, mixed response = 1, stable disease = 8; Ref. 73). A Phase II study proposal is under development by the Children’s Oncology Group for patients with newly diagnosed metastatic OS. The response rate of trimetrexate and simultaneous leucovorin would be predicted to be higher in newly diagnosed patients than in the setting of a conventional Phase II trial, because dihydrofolate reductase overexpression is not as frequent at diagnosis as compared with relapse. Furthermore, trimetrexate with leucovorin would be predicted to be at least as efficacious as high-dose methotrexate with less toxicity. The lack of renal toxicity may allow chemotherapy dose intensification, because it may be possible to administer trimetrexate simultaneously with other routine chemotherapy.

Microarrays and Preclinical OS Models.
With the goal of developing new therapeutic agents, several approaches have been taken to identify new therapeutic targets and to develop model systems for the evaluation of new agents. These model systems include cDNA microarrays, mouse OS syngeneic models, OS murine xenografts, and spontaneous OS in canines. Each of these systems will be discussed in the following sections.

Gene and Protein Expression.
cDNA microarray profiles of gene expression using patient tumor samples are now being used as a resource to catalogue candidate drug targets (74). Distinct cancers have distinct patterns of gene expression, and robust formal diagnostic classifications can use gene expression profiles to define novel subgroups of cancers that do not have previously defined clinicopathological correlates. As this information is compiled, the clinical application of the array data may require only a small group of genes to determine the tumor diagnosis or clinical subgroup. This information will be used to improve diagnostic categorization of tumors, to provide useful prognostic markers for outcome or likely therapeutic response, and to identify critical cancer cell pathways that can act as novel drug targets. Data analysis from a pilot study of bone tumors (ESFT and OS) demonstrates a strong cluster of ESFT-specific genes, a larger group of OS-associated genes, and a very strong difference in gene expression between the two tumors types. Whereas the OS samples demonstrated significant variability, the expression profiles of coupled primary and metastatic OS tumors were very similar. Additional analytic comparison of these tumor groups is under way to detect genes that are up-regulated or down-regulated during the metastatic process and to compare patient samples from children whose OS tumor demonstrated a good or poor response to induction chemotherapy. Lists of genes associated with survival are being generated, but the interpretation of these gene lists is just beginning. Biostatistical methods to evaluate the significance of gene expression within a cellular pathway will be needed to analyze the data being generated (75).

The expression of genes associated with the "metastatic phenotype" may be predictive of outcome in OS, given that metastatic clinical events are the proximate cause of OS-related death. However, only a small proportion of tumor cells from the gross tumor actually undergo metastasis, and tumor heterogeneity may mislead the metastatic phenotype evaluation obtained from small biopsy samples. The c-met proto-oncogene is of interest as the binding of c-met ligand can activate the tumor cell growth cycle, and start pathways involved in cell motility and the lysis of basement membranes, all components of the metastatic process. Tissue microarrays are being used to examine c-met expression levels in OS samples. Preliminary studies indicate that ~60% of OS tissue samples express c-met (76, 77). The matrix metalloproteinases are a family of proteins involved in the lysis and remodeling of the extracellular matrix in the metastatic process of OS. Expression of metalloproteinases 2, 9, and 14 has been associated with OS tumor growth and invasion.

Investigators are also looking at protein expression to monitor the molecules that directly control cellular activities. Whereas mRNA expression does not necessarily correlate with protein expression, the technical difficulty in directly monitoring protein expression makes mRNA expression profiling, comparative genomic hybridization for chromosomal gains/losses/amplifications, and spectral karyotyping to detect genetic translocations more popular currently (78, 79).25 Information from patient samples is being used to compare mRNA expression in primary OS versus normal osteoblasts, OS before and after chemotherapy, primary OS versus lung metastases, to predict response to neoadjuvant therapy, and as a prognostic factor.

Preclinical Models of OS.
The preclinical evaluation of new agents in OS is carried out using spontaneous, syngeneic, and xenogeneic or orthotopic animal models (80). As an example, the evaluation of camptothecin, epothilones, and signaling inhibitors has been performed in OS xenograft models (81). Validation issues for these OS models include the ability of the model to prospectively identify clinically active agents, the clinical heterogeneity of the model, and the false prediction rate of the model. Some relative correlation does exist between the xenograft models and clinical experience. Irinotecan has some activity in OS xenografts, but OS is much less sensitive than rhabdomyosarcoma, neuroblastoma, or Wilms tumor models. Epothilone B has some activity, but the signaling inhibitors imitanib mesylate (Gleevec) and gefitinib (Iressa) do not have significant single agent activity.

The genetic study of OS tumor cell metastatic potential has included a syngeneic murine OS model characterized by clonally related variants (K7M2 and K12) that differ in metastatic potential (82, 83). cDNA microarray analysis has defined 59 differentially expressed genes in a comparison of K7M2 and K12 primary tumors. K7M2 overexpressed genes related to cytoskeletal motility, adherence, and angiogenesis in comparison with K12, including the cytoskeleton linker protein Ezrin. A tissue microarray developed from canine OS tumors demonstrated that metastases have a statistically higher level of Ezrin expression compared with primary tumors. Ezrin mutants and the regulation of the protein during metastasis are under evaluation in murine models, and phosphorylated Ezrin expression is being evaluated in canine OS.

Canine OS provides an in vivo model with parallels to human OS (84, 85, 86). OS occurs spontaneously primarily in large-size breeds (most commonly in purebred dogs, with a male preponderance) affecting the long bones. The majority of canine OS cases are stage 2b, according to the Enneking staging system (87), at presentation and of high-grade histology. Approximately 15% are stage 3 at presentation, and pulmonary metastases are common. Local control is achieved by amputation or limb-sparing surgery. The occurrence of metastatic disease is greater in the lungs than bone, as in humans, but the time to local or distant recurrence is much shorter in dogs. The dog model provides a ready source for controlled preclinical treatment studies using a spontaneously arising tumor. Accrual of dogs to clinical study is rapid, and autopsy compliance is common. For example, a Colorado State University protocol examining the role of limb-sparing surgery, chemotherapy, and radiation included eligibility criteria of "localized" disease and <50% bone length involvement (88). Neoadjuvant therapy included arterial cisplatin on days 1 and 21, and radiotherapy followed by surgery. Increasing tumor necrosis (measured at surgery) correlated with a decreased local recurrence rate. Cisplatin and radiation increased the percentage of necrosis over radiation alone. Interestingly, dogs with infected limb repairs lived twice as long as dogs without infection. A study of the IGF-I inhibitor, OncoLar, randomized dogs with OS to OncoLar treatment in addition to chemotherapy and amputation. Sixty-four dogs were accrued to the study in just 8 months, but there was no difference in primary tumor necrosis, tumor cell apoptosis, or survival between the two treatment arms (89). New treatment approaches now under investigation include gemcitabine, antiangiogenic agents, gene therapy immunotherapy, and bisphosphonates. Basic biology studies are ongoing to examine the influence of IGF-I, and growth hormone in OS tumor growth and metastasis, and to evaluate growth hormone and growth hormone-releasing hormone antagonists.

New Treatment Strategies
Growth Hormone and IGF-I.
A series of clinical and preclinical observations may link IGF-I and OS. Adolescence is the age of the peak incidence of human OS, and the peak period of physiological growth hormone and IGF-I circulating concentrations. The prevalence of OS is greatest in dog breeds with the highest IGF-I concentrations. Furthermore, hypophysectomy is associated with antitumor activity in rodent OS due to IGF-I inhibition (90), and growth hormone blockade results in IGF-I inhibition and decreased OS growth. OS cell lines, which express IGF-I receptors, are IGF-I dependent for in vitro growth and survival (91). Although the circulating levels of IGF-I and IGFBP-3 among a few patients were not found to be predictive of the incidence or clinical behaviors of OS, a large prospective study of IGF-I and IGFBP-3 circulating levels is under way in OS patient serum. IGF-I levels have been investigated in ESFT using patient samples. Preliminary results demonstrate that IGF-I and IGFBP-3 levels can identify patients with the most widespread disease, but these levels are not independent predictors of prognosis. A relationship between the ratio of IGF-I and IGFBP-3 levels and prognosis has also been found (92). Blockade of IGF-I has been investigated using the sustained release sandostatin OncoLar, which can block IGF-I activity. A Phase I study of OncoLar with and without tamoxifen in relapsed OS patients demonstrated that OncoLar treatment leads to a sustained 40–50% decrease in IGF-I levels. The growth hormone level was not affected by OncoLar, and tamoxifen had no impact on measured IGF-I levels (93). There was no measurable tumor response to the OncoLar treatment. Investigators are considering combining OncoLar treatment with conventional chemotherapy hoping that a synergistic effect on tumor apoptosis may result (93).

Samarium.
Radiation therapy by either external beam radiation (94, 95) and/or a bone-seeking isotope can be effective in OS (96, 97). Samarium is a bone-seeking radiopharmaceutical that provides therapeutic irradiation to osteoblastic bone metastases. A recent clinical study administered samarium by a 30-min central line infusion to patients with bone metastases and dosimetry was performed on days 1, 2, and 5, to estimate tumor dose and residual radioactivity (96). Peripheral blood stem cell infusion on day 14 is necessary for patients receiving >3 mCi/kg, with day 13 radioactivity estimated at <3.6 mCi. Nonhematological acute toxicity includes manageable hypocalcemia during the infusion, and a "flare" reaction of bone pain within a day of samarium infusion was common, but pain control improved in all of the patients after samarium treatment.

The use of a radiation sensitizer, gemcitabine (98, 99, 100), after the samarium dose may increase the radiobiological effectiveness of this treatment. A trial using high-dose samarium day 0, gemcitabine day +1, with peripheral blood stem cell infusion day +14 is under way at Mayo Clinic. Additional means to increase bone formation and, thus, attract bone-seeking radioisotope could additionally improve the therapeutic index (101). The most effective use of samarium for control of OS bone lesions will probably be in combination with external beam radiotherapy.

Antibody-Based Strategies for Metastatic Solid Tumors.
Murine and human-mouse chimeric antibodies are under clinical investigation as directly targeted immunotherapy and anti-idiotype vaccines. The GD2 disialoganglioside is a target antigen expressed on OS and neuroblastoma cells, and GD2 is recognized by the 3F8, 5F11, and ch14.18 monoclonal antibodies. Early clinical studies of these monoclonal antibodies included patients with relapsed OS and neuroblastoma, and a randomized Phase III study of the anti-GD2 antibody (ch14.18) in high-risk neuroblastoma patients postautologous stem cell transplant is currently underway in the Children’s Oncology Group. Similarly, gp58 is a surface glycoprotein found on many pediatric tumors including OS, and clinical studies of the gp58-specific 8H9 monoclonal antibody are under way (102). In addition, ß-glucan, a glucose polymer extracted from plants and yeasts, has been studied in murine tumor models and found to significantly enhance the antitumor effect of mouse or chimeric monoclonal antibodies, prolonging progression-free survival (103). ß-Glucans are thought to bind to the lectin site of phagocytic and natural killer cells, thus facilitating antibody-dependent and complement-mediated tumor cell killing triggered by monoclonal antibodies. Finally, antibody vaccines designed to induce anti-idiotype anti-GD2 antibody are under clinical evaluation in children with relapsed neuroblastoma and may have applicability to OS as well.

Monoclonal antibodies have also been investigated in a pretargeting strategy using scFv-streptavidin to target the antibody fragment and therapeutic agent (radioisotopes and biologics) at the tumor. After tumor targeting of scFv-streptavidin, a clearing agent containing biotin and N-acetyl-galactosamine removes residual fusion proteins from the blood, significantly improving the tumor:blood drug exposure ratio. Using GD2 as a target, whole antibody or antibody-fragment alone results in a tumor:blood drug exposure ratio of <5:1, whereas the pretargeting approach gives a >50:1 ratio. This targeting and clearing approach focuses a high level of active agent to the tumor and diminishes nonspecific systemic exposure. T cells can also be targeted by modifying the T cells using chimeric immune receptors (T-bodies). The expanded, modified T cells recognize a predetermined antigen and proliferate when an anti-idiotype is added (104). The applicability of these targeted approaches in OS remains unclear.

Gene Therapy.
Gene therapy targeting an osteocalcin promoter has been explored in early phase clinical trials of prostate cancer. Osteocalcin is a noncollagenous protein produced by osteoblasts during bone mineralization. Immunohistochemical staining demonstrates osteocalcin production in osteoblastic OS (100%) and fibroblastic OS (70%). An adenoviral vector, containing the toxic TK gene, has been designed to regulate TK expression via the osteocalcin promoter (105). Transfected tumor cells are rendered sensitive to the acyclovir nucleotide analogue by the acquired TK activity, and the tumor cells die on administration of acyclovir.

Preclinical studies using the Ad-OC-TK/ACV therapy have demonstrated in vitro and in vivo growth inhibition for rat OS, including localized and metastatic OS in vivo models (106, 107). Phase I studies of the AD-OC-TK vector delivered by intratumoral injection in prostate patients demonstrated good tolerability of the injection and no unexpected toxicities (108).

Inhaled GM-CSF.
Preclinical studies of melanoma in mice have demonstrated the ability of GM-CSF-transfected tumor cells to induce potent, specific, long-lasting immunity to subsequent tumor challenge. Tumors transduced by GM-CSF demonstrate dense macrophage infiltration, and locally produced GM-CSF facilitates killing of nontransduced bystander tumor cells. Aerosolized GM-CSF has been investigated for delivery of the cytokine to lung tissue affected by metastatic tumor. The aerosolized cytokine is absorbed by pulmonary lymphatics and can interact with receptor-bearing cells in bronchial and pulmonary lymph nodes.

A Phase I study of aerosolized GM-CSF in solid tumor patients at the Mayo Clinic allowed intrapatient dose escalation from 60 to 240 µg/inhalation with twice daily administration for 1 week and 1-week rest (109). Among the 6 patients with stable disease, 5 continued on GM-CSF for 2–8 months after reaching the third dose level without significant toxicity. A subsequent Phase II study of 45 patients included 8 OS patients. Two of 8 OS patients with gross disease progressed, whereas the other 6 achieved a complete response with surgery, and 2 have been disease-free for 2.5 and 4 years. Clinical studies under consideration include administration of aerosolized GM-CSF before thoracotomy in OS patients with isolated pulmonary relapses. For relapsed patients with bilateral pulmonary nodules, a staged thoracotomy would allow examination of pulmonary nodules resected before and after treatment with GM-CSF to determine the possible biological and pathological changes associated with the aerosolized GM-CSF treatment. Given the demonstrated lack of significant toxicity, a pilot study of aerosolized GM-CSF for 6 months postsurgery in patients with nonmetastatic OS has been proposed.

Localized Platinum.
Veterinary investigators have developed an absorbable sponge-like device to deliver slow-release cisplatin in the surgical site (110, 111). This localized treatment approach can address: (a) local recurrence; (b) marginal resections; (c) tumor close to the wound bed; and (d) the need for a localized tumor dose of platinum with minimal systemic exposure. The slow-release platinum also appears to have a radiosensitizing activity (112). A study in 80 dogs that underwent OS surgery resection and allograft placement included i.v. cisplatin therapy for all of the dogs, with a randomization to the platinum implant device. The local recurrence rate was 15% among the implant cohort compared with 55% for the control group. Among those dogs with histologically incomplete tumor resection, a statistically significant decrease in the recurrence rate was found.

Bisphosphonates.
Bisphosphonates bind strongly to hydroxyapatite on the bone surface, have direct inhibitory effects on osteoclast-mediated bone resorption, and affect osteoblast activity (113). Whereas no direct cytotoxic effect of bisphosphonates is reported, the agents can induce bone cell apoptosis, inhibit cytokine production (Ref. 114; IL-6 promotes OS cell growth and encourages angiogenesis), and inhibit metalloproteinases. The bisphosphonates can prevent bone loss and fractures, and the prevention of bone metastasis development has been demonstrated in breast cancer patients with microscopic bone marrow "metastases" (115, 116, 117). The bisphosphonates may have several potential roles in sarcoma treatment, including the prevention of: tumor metastases, osteoporosis associated with chemotherapy (118), and restricted weight bearing, periprosthetic osteolysis, and bone loss from the stress bypass effect (119, 120, 121). Although the agents have been used to treat osteogenesis imperfecta in children (114), pediatric data are limited, and the role of bisphosphonates in the prevention of metastatic tumor development or primary tumor progression remains to be defined.

Angiogenesis and Metronomic Low-Dose Chemotherapy.
The use of low-dose chemotherapy to block tumor-associated endothelial cell proliferation has gained interest as an antiangiogenic approach to tumor treatment. Using a tolerable, chronic low dose of select chemotherapy agents, investigators believe that endothelial cell proliferation can be blocked (122) without the acquisition of drug resistance (the target cells are genetically stable). COX-2 inhibitors have been added to the chemotherapy regimen due to the expression of COX-2 in tumor neovasculature, neoplastic cells, and stromal cells (123). Mice lacking COX-2 expression demonstrate that decreased VEGF expression and tumors in such mice grow more slowly in association with reduced tumor angiogenesis. Additionally, COX-2 inhibitors decrease VEGF production and VEGF-induced endothelial cell activation. An ongoing clinical pilot study is evaluating the safety and toxicity of chronically administering celecoxib and low-dose vinblastine or cyclophosphamide in pediatric patients with recurrent solid tumors (124). As part of the current pilot study, the pharmacokinetics of celecoxib in children is being studied (125), and angiogenic growth factors (VEGF, basic fibroblast growth factor, and vascular cell adhesion molecule 1) are being evaluated as surrogate markers of angiogenesis. Dynamic magnetic resonance imaging is also being assessed as a tool for specific antiangiogenic tumor response. The treatment regimen has been well tolerated, and the surrogate markers have demonstrated a large interpatient variability (125) .

Conclusion

Investigators are pursuing a greater understanding of OS biology at the subcellular level by developing gene and protein expression array data that may soon provide customized information on tumor prognosis and metastatic potential, as well as indications of possible tumor targets for selective therapy. The broader employment of animal models, and especially spontaneous canine OS, can provide a much-needed means to study the activity of new treatment interventions at the preclinical level. Finally, a variety of local and targeted therapies are now under evaluation in the clinic. With these new treatments, investigators can envision that targeted agents, such as new small molecule inhibitors, growth factors, radionucleotides, viral gene therapy, and various monoclonal antibodies, may one day complement the systemic chemotherapy and surgery that remain the foundation of OS treatment.

Footnotes

23 The abbreviations used are: OS, osteosarcoma; ESFT, Ewing Sarcoma Family Tumor; Rb, retinoblastoma; LOH, loss of heterozygosity; PTH, parathyroid hormone; ALT, alternative telomere lengthening; FasL, Fas ligand; IL, interleukin; Ad.mIL-12, adenoviral murine interleukin 12; MTP-PE, muramyl tripeptide phosphatidylethanolamine; MDR1, multidrug resistance gene 1; IGF, insulin-like growth factor; IGFBP-3, insulin-like growth factor binding protein 3; TK, thymidine kinase; GM-CSF, granulocyte macrophage colony-stimulating factor; COX, cyclooxygenase; VEGF, vascular endothelial growth factor.

24 A. A. Sandberg and J. A. Bridge. Updates on the cytogenetics and molecular genetics of bone and soft tissue tumors: osteosarcoma and related tumors. Cancer Genetics Cytogenetics, submitted, 2003.

25 C. C. Lau, C. P. Harris, X-Y. Lu, L. Perlaky, S. Gogineni, M. Chintagumpala, J. Hicks, A. G. Huvos, P. A. Meyers, J. H. Healey, R. Gorlick, and P. H. Rao. Frequent amplification and rearrangement of chromosomal bands 6p12-p21 and 17p11.2 in osteosarcomas. Mol. Cancer Ther., submitted, 2003.

Received 5/1/03; revised 7/29/03; accepted 7/29/03.

References

  1. Mitelman F. Recurrent chromosome aberrations in cancer. Mutat. Res., 462: 247-253, 2000.[Medline]
  2. Bridge J. A., Nelson M., McComb E., McGuire M. H., Rosenthal H., Vergara G., Maale G. E., Spanier S., Neff J. R. Cytogenetic findings in 73 osteosarcoma specimens and a review of the literature. Cancer Genetics Cytogenetics, 95: 74-87, 1997.
  3. Boehm A. K., Squire J. A., Bayani J., Nelson M., Neff J. R., Bridge J. A. Cytogenetic findings in 36 osteosarcoma specimens and a review of the literature. Ped. Pathol. Mol. Med., 19: 359-376, 2000.
  4. Gisselsson D., Palsson E., Hoglund M., Domanski H., Mertens F., Pandis N., Sciot R., Dal Cin P., Bridge J. A., Mandahl N. Differentially amplified chromosome 12 sequences in low- and high-grade osteosarcoma. Genes Chromosomes Cancer, 33: 133-140, 2002.[Medline]
  5. Raymond, A. K., Ayala, A. G., and Knuutila, S. Conventional Osteosarcoma. In: P. Kleihues, L. Sobin, C. Fletcher, K. Unni, and F. Mertens (eds.), WHO Classification of Tumours: Pathology and Genetics of Tumours of Soft Tissue and Bone, pp. 267–269. Lyon, France: IARC Press, 2002.
  6. Bayani J., Zielenska M., Pandita A., Al-Romaih K., Karaskova J., Harrison K., Bridge J. A., Sorensen P., Thorner P., Squire J. A. SKY identifies recurrent complex rearrangements involving chromosomes 8, 17, and 20 in osteosarcomas. Genes Chromosome Cancer, 36: 7-16, 2003.[Medline]
  7. Ladanyi M., Gorlick R. Molecular pathology and molecular pharmacology of osteosarcoma. Pediatr. Pathol. Mol. Med., 19: 391-413, 2000.
  8. Kruzelock R. P., Hansen M. F. Molecular genetics and cytogenetics of sarcomas. Hematology - Oncology Clin. N. Am., 9: 513-540, 1995.
  9. Kruzelock R. P., Murphy E. C., Strong L. C., Naylor S. L., Hansen M. F. Localization of a novel tumor suppressor locus on human chromosome 3q important in osteosarcoma tumorigenesis. Cancer Res., 57: 106-109, 1997.[Abstract]
  10. Nellissery M. J., Padalecki S. S., Brkanac Z., Singer F. R., Roodman G. D., Unni K. K., Leach R. J., Hansen M. F. Evidence for a novel osteosarcoma tumor-suppressor gene in the chromosome 18 region genetically linked with Paget disease of bone. Am. J. Hum. Genet., 63: 817-824, 1998.[Medline]
  11. Malkin D., Li F. P., Strong L. C., Fraumeni J. F., Jr., Nelson C. E., Kim D. H., Kassel J., Gryka M. A., Bischoff F. Z., Tainsky M. A., et al Germ line p53 mutations in a familial syndrome of breast cancer, sarcomas, and other neoplasms. Science (Wash. DC), 250: 1233-1238, 1990.[Medline]
  12. Quesnel S., Malkin D. Genetic predisposition to cancer and familial cancer syndromes. Ped. Clin. N. Am., 44: 791-808, 1997.[Medline]
  13. Malkin D. The role of p53 in human cancer. J. Neuro-Oncol., 51: 231-243, 2001.[Medline]
  14. Wei G., Lonardo F., Ueda T., Kim T., Huvos A. G., Healey J. H., Ladanyi M. CDK4 gene amplification in osteosarcoma: reciprocal relationship with INK4A gene alterations and mapping of 12q13 amplicons. Int. J. Cancer, 80: 199-204, 1999.[Medline]
  15. Gokgoz N., Wunder J. S., Mousses S., Eskandarian S., Bell R. S., Andrulis I. L. Comparison of p53 mutations in patients with localized osteosarcoma and metastatic osteosarcoma. Cancer (Phila.), 92: 2181-2189, 2001.
  16. Yamaguchi T., Toguchida J., Yamamuro T., Kotoura Y., Takada N., Kawaguchi N., Kaneko Y., Nakamura Y., Sasaki M. S., Ishizaki K. Allelotype analysis in osteosarcomas: frequent allele loss on 3q, 13q, 17p, and 18q. Cancer Res., 52: 2419-2423, 1992.[Abstract]
  17. Tarkkanen M., Karhu R., Kallioniemi A., Elomaa I., Kivioja A. H., Nevalainen J., Bohling T., Karaharju E., Hyytinen E., Knuutila S., et al Gains and losses of DNA sequences in osteosarcomas by comparative genomic hybridization. Cancer Res., 55: 1334-1338, 1995.[Abstract]
  18. Hoogerwerf W. A., Hawkins A. L., Perlman E. J., Griffin C. A. Chromosome analysis of nine osteosarcomas. Genes, Chromosomes Cancer, 9: 88-92, 1994.[Medline]
  19. Bianco P., Riminucci M., Gronthos S., Robey P. Bone marrow stromal cells: nature, biology and potential applications. Stem Cells, 19: 180-192, 2001.[Abstract/Free Full Text]
  20. Ohyama K., Chung C. H., Chen E., Gibson C. W., Misof K., Fratzl P., Shapiro I. M. p53 influences mice skeletal development. J.Craniofacial Genet. Dev. Biol., 17: 161-171, 1997.
  21. Vahle J., Sato M., Long G., Young J., Francis P., Engelhardt J., Westmore M., Ma Y., Nold J. Skeletal changes in rats given daily subcutaneous injections of recombinant human parathyroid hormone (1–34) for 2 years and relevance to human safety. Toxicol. Pathol., 30: 312-321, 2002.[Medline]
  22. Milas J., Stanislaus D., Ohashi N., Pulcini J., Murthy S., Leyvand I., Dunn-Jena P., Hock J. In vivo, p53 contributes to the anabolic effect of PTH by controlling the magnitude of AP-1 gene expression. Ann. Mtg. Endocr. Soc., : 2001.
  23. Stanislaus D., Yang X., Liang J., Wolfe J., Cain R., Onyia J., Falla N., Marder P., Bidwell J., Queener S., Hock J. In vivo regulation of apoptosis in metaphyseal trabecular bone of young rats by synthetic human parathyroid hormone, hPTH 1–34 fragment. Bone, 27: 209-218, 2000.[Medline]
  24. Artandi S. E., Chang S., Lee S. L., Alson S., Gottlieb G. J., Chin L., DePinho R. A. Telomere dysfunction promotes non-reciprocal translocations and epithelial cancers in mice. Nature (Lond.), 406: 641-645, 2000.[Medline]
  25. Harley C. B. Telomere loss: mitotic clock or genetic time bomb?. Mutat. Res., 256: 271-282, 1991.[Medline]
  26. Bodnar A. G., Ouellette M., Frolkis M., Holt S. E., Chiu C. P., Morin G. B., Harley C. B., Shay J. W., Lichtsteiner S., Wright W. E. Extension of life-span by introduction of telomerase into normal human cells.[comment]. Science (Wash. DC), 279: 349-352, 1998.[Abstract/Free Full Text]
  27. Shay J. W., Bacchetti S. A survey of telomerase activity in human cancer. Eur. J. Cancer, 33: 787-791, 1997.[Medline]
  28. Bryan T. M., Englezou A., Dalla-Pozza L., Dunham M. A., Reddel R. R. Evidence for an alternative mechanism for maintaining telomere length in human tumors and tumor-derived cell lines. Nat. Med., 3: 1271-1274, 1997.[Medline]
  29. Scheel C., Schaefer K. L., Jauch A., Keller M., Wai D., Brinkschmidt C., van Valen F., Boecker W., Dockhorn-Dworniczak B., Poremba C. Alternative lengthening of telomeres is associated with chromosomal instability in osteosarcomas. Oncogene, 20: 3835-3844, 2001.[Medline]
  30. Aue G., Muralidhar B., Schwartz H. S., Butler M. G. Telomerase activity in skeletal sarcomas. Ann. Surg. Oncol., 5: 627-634, 1998.[Abstract]
  31. Sangiorgi L., Gobbi G. A., Lucarelli E., Sartorio S. M., Mordenti M., Ghedini I., Maini V., Scrimieri F., Reggiani M., Bertoja A. Z., Benassi M. S., Picci P. Presence of telomerase activity in different musculoskeletal tumor histotypes and correlation with aggressiveness. Int. J. Cancer, 95: 156-161, 2001.[Medline]
  32. Chang S., Khoo C. M., Naylor M. L., Maser R. S., DePinho R. A. Telomere-based crisis: functional differences between telomerase activation and ALT in tumor progression (Published erratum appears in Genes Dev., 17: 541, 2003]. Genes Dev., 17: 88-100, 2003.[Abstract/Free Full Text]
  33. Carbone M., Rizzo P., Procopio A., Giuliano M., Pass H. I., Gebhardt M. C., Mangham C., Hansen M., Malkin D. F., Bushart G., Pompetti F., Picci P., Levine A. S., Bergsagel J. D., Garcea R. L. SV40-like sequences in human bone tumors. Oncogene, 13: 527-535, 1996.[Medline]
  34. Mendoza S. M., Konishi T., Miller C. W. Integration of SV40 in human osteosarcoma DNA. Oncogene, 17: 2457-2462, 1998.[Medline]
  35. Lednicky J. A., Stewart A. R., Jenkins J. J., 3rd, Finegold M. J., Butel J. S. SV40 DNA in human osteosarcomas shows sequence variation among T-antigen genes. Int. J. Cancer, 72: 791-800, 1997.[Medline]
  36. Malkin D., Chilton-MacNeill S., Meister L. A., Sexsmith E., Diller L., Garcea R. L. Tissue-specific expression of SV40 in tumors associated with the Li-Fraumeni syndrome. Oncogene, 20: 4441-4449, 2001.[Medline]
  37. Lefevre S. H., Vogt N., Dutrillaux A. M., Chauveinc L., Stoppa-Lyonnet D., Doz F., Desjardins L., Dutrillaux B., Chevillard S., Malfoy B. Genome instability in secondary solid tumors developing after radiotherapy of bilateral retinoblastoma. Oncogene, 20: 8092-8099, 2001.[Medline]
  38. Poulaki V., Mitsiades N., Romero M. E., Tsokos M. Fas-mediated apoptosis in neuroblastoma requires mitochondrial activation and is inhibited by FLICE inhibitor protein and Bcl-2. Cancer Res., 61: 4864-4872, 2001.[Abstract/Free Full Text]
  39. Mitsiades N., Poulaki V., Mitsiades C., Tsokos M. Ewing’s sarcoma family tumors are sensitive to tumor necrosis factor-related apoptosis-inducing ligand and express death receptor 4 and death receptor 5. Cancer Res., 61: 2704-2712, 2001.[Abstract/Free Full Text]
  40. Hewitt R. E., McMarlin A., Kleiner D., Wersto R., Martin P., Tsokos M., Stamp G. W., Stetler-Stevenson W. G., Tsoskas M. Validation of a model of colon cancer progression (Published erratum appears in J. Pathol. 194: 507, 2001). J. Pathol., 192: 446-454, 2000.[Medline]
  41. Mitsiades N., Poulaki V., Mastorakos G., Tseleni-Balafouta S. T., Kotoula V., Koutras D. A., Tsokos M. Fas ligand expression in thyroid carcinomas: a potential mechanism of immune evasion. J. Clin. Endocrinol. Metab., 84: 2924-2932, 1999.[Abstract/Free Full Text]
  42. Mitsiades N., Poulaki V., Leone A., Tsokos M. Fas-mediated apoptosis in Ewing’s sarcoma cell lines by metalloproteinase inhibitors. J. Natl. Cancer Instit., 91: 1678-1684, 1999.[Abstract/Free Full Text]
  43. Worth L. L., Lafleur E. A., Jia S. F., Kleinerman E. S. Fas expression inversely correlates with metastatic potential in osteosarcoma cells. Oncol. Rep., 9: 823-827, 2002.[Medline]
  44. Jia S. F., Worth L. L., Kleinerman E. S. A nude mouse model of human osteosarcoma lung metastases for evaluating new therapeutic strategies. Clin. Exp. Metastasis, 17: 501-506, 1999.[Medline]
  45. Worth L. L., Jia S. F., Zhou Z., Chen L., Kleinerman E. S. Intranasal therapy with an adenoviral vector containing the murine interleukin-12 gene eradicates osteosarcoma lung metastases. Clin. Cancer Res., 6: 3713-3718, 2000.[Abstract/Free Full Text]
  46. Lafleur E. A., Jia S. F., Worth L. L., Zhou Z., Owen-Schaub L. B., Kleinerman E. S. Interleukin (IL)-12 and IL-12 gene transfer up-regulate Fas expression in human osteosarcoma and breast cancer cells. Cancer Res., 61: 4066-4071, 2001.[Abstract/Free Full Text]
  47. Jia S. F., Worth L. L., Densmore C. L., Xu B., Zhou Z., Kleinerman E. S. Eradication of osteosarcoma lung metastases following intranasal interleukin-12 gene therapy using a nonviral polyethylenimine vector. Cancer Gene Ther., 9: 260-266, 2002.[Medline]
  48. Meyers P., Gorlick R. Osteosarcoma. Ped. Clin. N. Am., 44: 973-989, 1997.[Medline]
  49. Wigginton J. M., Wiltrout R. H. IL-12/IL-2 combination cytokine therapy for solid tumours: translation from bench to bedside. Exp. Opin. Biol. Ther., 2: 513-524, 2002.
  50. Wigginton J. M., Komschlies K. L., Back T. C., Franco J. L., Brunda M. J., Wiltrout R. H. Administration of interleukin 12 with pulse interleukin 2 and the rapid and complete eradication of murine renal carcinoma. J. Natl. Cancer Inst., 88: 38-43, 1996.[Abstract/Free Full Text]
  51. Wigginton J. M., Park J. W., Gruys M. E., Young H. A., Jorcyk C. L., Back T. C., Brunda M. J., Strieter R. M., Ward J., Green J. E., Wiltrout R. H. Complete regression of established spontaneous mammary carcinoma and the therapeutic prevention of genetically programmed neoplastic transition by IL-12/pulse IL-2: induction of local T cell infiltration. Fas/Fas ligand gene expression, and mammary epithelial apoptosis. J. Immunol., 166: 1156-1168, 2001.[Abstract/Free Full Text]
  52. Wigginton J. M., Gruys E., Geiselhart L., Subleski J., Komschlies K. L., Park J. W., Wiltrout T. A., Nagashima K., Back T. C., Wiltrout R. H. IFN-γ and Fas/FasL are required for the antitumor and antiangiogenic effects of IL-12/pulse IL-2 therapy, J. Clin. Investig., 108: 51-62, 2001.
  53. Meyers P. A., Gorlick R. Osteosarcoma. Ped. Clin. N. Am., 44: 973-989, 1997.[Medline]
  54. Takeshita H., Kusuzaki K., Ashihara T., Gebhardt M. C., Mankin H. J., Hirasawa Y. Intrinsic resistance to chemotherapeutic agents in murine osteosarcoma cells. J. Bone Joint Surg. Am. Vol., 82-A: 963-969, 2000.[Abstract/Free Full Text]
  55. Baldini N., Scotlandi K., Barbanti-Brodano G., Manara M. C., Maurici D., Bacci G., Bertoni F., Picci P., Sottili S., Campanacci M., et al Expression of P-glycoprotein in high-grade osteosarcomas in relation to clinical outcome[comment]. N. Eng. J. Med., 333: 1380-1385, 1995.[Abstract/Free Full Text]
  56. Takeshita H., Gebhardt M. C., Springfield D. S., Kusuzaki K., Mankin H. J. Experimental models for the study of drug resistance in osteosarcoma: P-glycoprotein-positive, murine osteosarcoma cell lines. J. Bone Joint Surg. Am. Vol., 78: 366-375, 1996.[Abstract/Free Full Text]
  57. Baldini N., Scotlandi K., Serra M., Kusuzaki K., Shikita T., Manara M. C., Maurici D., Campanacci M. Adriamycin binding assay: a valuable chemosensitivity test in human osteosarcoma. J. Cancer Res. Clin. Oncol., 119: 121-126, 1992.[Medline]
  58. Wunder J. S., Bell R. S., Wold L., Andrulis I. L. Expression of the multidrug resistance gene in osteosarcoma: a pilot study. J. Orthopaed. Res., 11: 396-403, 1993.[Medline]
  59. Baldini N., Scotlandi K., Serra M., Picci P., Bacci G., Sottili S., Campanacci M. P-glycoprotein expression in osteosarcoma: a basis for risk-adapted adjuvant chemotherapy. J. Orthopaed. Res., 17: 629-632, 1999.[Medline]
  60. Stein U., Wunderlich V., Haensch W., Schmidt-Peter P. Expression of the mdr1 gene in bone and soft tissue sarcomas of adult patients. Eur. J. Cancer, 29A: 1979-1981, 1993.
  61. Posl M., Amling M., Grahl K., Hentz M., Ritzel H., Werner M., Winkler K., Delling G. P-glycoprotein expression in high grade central osteosarcoma and normal bone cells. An immunohistochemical study. Gen. Diag. Pathol., 142: 317-325, 1997.
  62. Chan H. S., Grogan T. M., Haddad G., DeBoer G., Ling V. P-glycoprotein expression: critical determinant in the response to osteosarcoma chemotherapy. J. Natl. Cancer Inst., 89: 1706-1715, 1997.[Abstract/Free Full Text]
  63. Wunder J. S., Bull S. B., Aneliunas V., Lee P. D., Davis A. M., Beauchamp C. P., Conrad E. U., Grimer R. J., Healey J. H., Rock M. J., Bell R. S., Andrulis I. L. MDR1 gene expression and outcome in osteosarcoma: a prospective, multicenter study. J. Clin. Oncol., 18: 2685-2694, 2000.[Abstract/Free Full Text]
  64. Bodey B., Taylor C. R., Siegel S. E., Kaiser H. E. Immunocytochemical observation of multidrug resistance (MDR) p170 glycoprotein expression in human osteosarcoma cells. The clinical significance of MDR protein overexpression. Anticancer Res., 15: 2461-2468, 1995.[Medline]
  65. Schwartz C. L., Gorlick R. G., Teot L. A., Grier H. E., Krailo M., Meyers P. A. A prospective study of p-glycoprotein (P-GP) expression in newly diagnosed non-metastatic osteosarcoma: An intergroup study of the Children’s Cancer Group and the Pediatric Oncology Group. Proc. Am. Soc. Clin. Oncol., 19: 587a 2000.
  66. Scotlandi K., Manara M. C., Serra M., Benini S., Maurici D., Caputo A., De Giovanni C., Lollini P. L., Nanni P., Picci P., Campanacci M., Baldini N. The expression of P-glycoprotein is causally related to a less aggressive phenotype in human osteosarcoma cells. Oncogene, 18: 739-746, 1999.[Medline]
  67. Bush J. A., Li G. Cancer chemoresistance: the relationship between p53 and multidrug transporters. Int. J. Cancer, 98: 323-330, 2002.[Medline]
  68. Banerjee D., Mayer-Kuckuk P., Capiaux G., Budak-Alpdogan T., Gorlick R., Bertino J. R. Novel aspects of resistance to drugs targeted to dihydrofolate reductase and thymidylate synthase. Biochim. Biophys. Acta., 1587: 164-173, 2002.[Medline]
  69. Guo W., Healey J. H., Meyers P. A., Ladanyi M., Huvos A. G., Bertino J. R., Gorlick R. Mechanisms of methotrexate resistance in osteosarcoma. Clin. Cancer Res., 5: 621-627, 1999.[Abstract/Free Full Text]
  70. Yang R., Sowers R., Mazza B., Healey J. H., Huvos A., Grier H., Bernstein M., Beardsley G. P., Krailo M. D., Devidas M., Bertino J. R., Meyers P. A., Gorlick R. Sequence alterations in the reduced folate carrier are observed in osteosarcoma tumor samples. Clin. Cancer Res., 9: 837-844, 2003.[Abstract/Free Full Text]
  71. Bacci G., Ferrari S., Delepine N., Bertoni F., Picci P., Mercuri M., Bacchini P., Brach del Prever A., Tienghi A., Comandone A., Campanacci M. Predictive factors of histologic response to primary chemotherapy in osteosarcoma of the extremity: study of 272 patients preoperatively treated with high-dose methotrexate, doxorubicin, and cisplatin[comment]. J. Clin. Oncol., 16: 658-663, 1998.[Abstract]
  72. Delepine N., Delepine G., Bacci G., Rosen G., Desbois J. C. Influence of methotrexate dose intensity on outcome of patients with high grade osteogenic osteosarcoma. Analysis of the literature. Cancer (Phila.), 78: 2127-2135, 1996.
  73. Trippett T., Meyers P., Gorlick R., Steinherz P., Wollner N., Bertino J. R. High dose trimetrexate with leucovorin protection in recurrent childhood malignancies: A Phase II trial. Proc. Am. Soc. Clin. Oncol., 18: 231a 1999.
  74. Wolf M., El-Rifai W., Tarkkanen M., Kononen J., Serra M., Eriksen E. F., Elomaa I., Kallioniemi A., Kallioniemi O. P., Knuutila S. Novel findings in gene expression detected in human osteosarcoma by cDNA microarray. Cancer Genet. Cytogenet., 123: 128-132, 2000.[Medline]
  75. Khan J., Wei J. S., Ringner M., Saal L. H., Ladanyi M., Westermann F., Berthold F., Schwab M., Antonescu C. R., Peterson C., Meltzer P. S. Classification and diagnostic prediction of cancers using gene expression profiling and artificial neural networks. Nat. Med., 7: 673-679, 2001.[Medline]
  76. Oda Y., Naka T., Takeshita M., Iwamoto Y., Tsuneyoshi M. Comparison of histological changes and changes in nm23 and c-MET expression between primary and metastatic sites in osteosarcoma: a clinicopathologic and immunohistochemical study. Hum. Pathol., 31: 709-716, 2000.[Medline]
  77. Ferracini R., Angelini P., Cagliero E., Linari A., Martano M., Wunder J., Buracco P. MET oncogene aberrant expression in canine osteosarcoma. J. Orthopaed. Res., 18: 253-256, 2000.[Medline]
  78. Zielenska M., Bayani J., Pandita A., Toledo S., Marrano P., Andrade J., Petrilli A., Thorner P., Sorensen P., Squire J. A. Comparative genomic hybridization analysis identifies gains of 1p35 approximately p36 and chromosome 19 in osteosarcoma. Cancer Genet. Cytogenet., 130: 14-21, 2001.[Medline]
  79. Ozaki T., Neumann T., Wai D., Schafer K. L., van Valen F., Lindner N., Scheel C., Bocker W., Winkelmann W., Dockhorn-Dworniczak B., Horst J., Poremba C. Chromosomal alterations in osteosarcoma cell lines revealed by comparative genomic hybridization and multicolor karyotyping. Cancer Genet. Cytogenet., 140: 145-152, 2003.[Medline]
  80. Houghton P. J., Adamson P. C., Blaney S., Fine H. A., Gorlick R., Haber M., Helman L., Hirschfeld S., Hollingshead M. G., Israel M. A., Lock R. B., Maris J. M., Merlino G., Patterson W., Reynolds C. P., Shannon K., Yu A., Yu J., Smith M. A. Testing of new agents in childhood cancer preclinical models: meeting summary. Clin. Cancer Res., 8: 3646-3657, 2002.[Abstract/Free Full Text]
  81. Thompson J., Stewart C. F., Houghton P. J. Animal models for studying the action of topoisomerase I targeted drugs. Biochim. Biophys. Acta., 1400: 301-319, 1998.[Medline]
  82. Khanna C., Prehn J., Yeung C., Caylor J., Tsokos M., Helman L. An orthotopic model of murine osteosarcoma with clonally related variants differing in pulmonary metastatic potential. Clin. Exp. Metastasis, 18: 261-271, 2000.[Medline]
  83. Khanna C., Khan J., Nguyen P., Prehn J., Caylor J., Yeung C., Trepel J., Meltzer P., Helman L. Metastasis-associated differences in gene expression in a murine model of osteosarcoma. Cancer Res., 61: 3750-3759, 2001.[Abstract/Free Full Text]
  84. Withrow S. J., Powers B. E., Straw R. C., Wilkins R. M. Comparative aspects of osteosarcoma. Dog versus man. Clin. Orthopaed. Rel. Res., 270: 159-68, 1991.[Medline]
  85. Withrow S. J. Osteosarcoma. Veterinary Quarterly, 20: S19-S21, 1998.[Medline]
  86. Mendoza S., Konishi T., Dernell W. S., Withrow S. J., Miller C. W. Status of the p53. Rb and MDM2 genes in canine osteosarcoma. Anticancer Res., 18: 4449-4453, 1998.[Medline]
  87. Enneking W., Spanier S., Goodman M. A system for the surgical staging of musculoskeletal sarcoma. Clin. Orthop., 153: 106-120, 1980.[Medline]
  88. Withrow S. J., Thrall D. E., Straw R. C., Powers B. E., Wrigley R. H., Larue S. M., Page R. L., Richardson D. C., Bissonette K. W., Betts C. W., et al Intra-arterial cisplatin with or without radiation in limb-sparing for canine osteosarcoma. Cancer (Phila.), 71: 2484-2490, 1993.
  89. Khanna C., Prehn J., Hayden D., Cassaday R. D., Caylor J., Jacob S., Bose S. M., Hong S. H., Hewitt S. M., Helman L. J. A randomized controlled trial of octreotide pamoate long-acting release and carboplatin versus carboplatin alone in dogs with naturally occurring osteosarcoma: evaluation of insulin-like growth factor suppression and chemotherapy. Clin. Cancer Res., 8: 2406-2412, 2002.[Abstract/Free Full Text]
  90. Pollak M., Sem A. W., Richard M., Tetenes E., Bell R. Inhibition of metastatic behavior of murine osteosarcoma by hypophysectomy. J. Natl. Cancer Inst., 84: 966-971, 1992.[Abstract]
  91. Raile K., Hoflich A., Kessler U., Yang Y., Pfuender M., Blum W. F., Kolb H., Schwarz H. P., Kiess W. Human osteosarcoma (U-2 OS) cells express both insulin-like growth factor-I (IGF-I) receptors and insulin-like growth factor-II/mannose-6-phosphate (IGF-II/M6P) receptors and synthesize IGF-II: autocrine growth stimulation by IGF-II via the IGF-I receptor. J. Cell. Physiol., 159: 531-541, 1994.[Medline]
  92. Toretsky J. A., Steinberg S. M., Thakar M., Counts D., Pironis B., Parente C., Eskenazi A., Helman L., Wexler L. H. Insulin-like growth factor type 1 (IGF-1) and IGF binding protein-3 in patients with Ewing sarcoma family of tumors. Cancer (Phila.), 92: 2941-2947, 2001.
  93. Mansky P. J., Liewehr D. J., Steinberg S. M., Chrousos G. P., Avila N. A., Long L., Bernstein D., Mackall C. L., Hawkins D. S., Helman L. J. Treatment of metastatic osteosarcoma with the somatostatin analog OncoLar: significant reduction of insulin-like growth factor-1 serum levels. J. Ped. Hematol./Oncol., 24: 440-446, 2002.
  94. Kamada T., Tsujii H., Tsuji H., Yanagi T., Mizoe J. E., Miyamoto T., Kato H., Yamada S., Morita S., Yoshikawa K., Kandatsu S., Tateishi A., Working Group for the Bone and Soft Tissue Sarcomas. Efficacy and safety of carbon ion radiotherapy in bone and soft tissue sarcomas. J. Clin. Oncol., 20: 4466-4471, 2002.[Abstract/Free Full Text]
  95. Machak G. N., Tkachev S. I., Solovyev Y. N., Sinyukov P. A., Ivanov S. M., Kochergina N. V., Ryjkov A. D., Tepliakov V. V., Bokhian B. Y., Glebovskaya V. V. Neoadjuvant chemotherapy and local radiotherapy for high-grade osteosarcoma of the extremities. Mayo Clinic Proc., 78: 147-155, 2003.[Medline]
  96. Anderson P. M., Wiseman G. A., Dispenzieri A., Arndt C. A., Hartmann L. C., Smithson W. A., Mullan B. P., Bruland O. S. High-dose samarium-153 ethylene diamine tetramethylene phosphonate: low toxicity of skeletal irradiation in patients with osteosarcoma and bone metastases. J. Clin. Oncol., 20: 189-196, 2002.[Abstract/Free Full Text]
  97. Anderson P. M. Radiotherapy is effective for osteosarcoma responding to chemotherapy (editorial). Mayo Clinic Proc., : 2003.
  98. Lawrence T. S., Eisbruch A., McGinn C. J., Fields M. T., Shewach D. S. Radiosensitization by gemcitabine. Oncology (Huntingt.), 13: 55-60, 1999.
  99. Eisbruch A., Shewach D. S., Bradford C. R., Littles J. F., Teknos T. N., Chepeha D. B., Marentette L. J., Terrell J. E., Hogikyan N. D., Dawson L. A., Urba S., Wolf G. T., Lawrence T. S. Radiation concurrent with gemcitabine for locally advanced head and neck cancer: a phase I trial and intracellular drug incorporation study. J. Clin. Oncol., 19: 792-799, 2001.[Abstract/Free Full Text]
  100. McGinn C. J., Lawrence T. S. Recent advances in the use of radiosensitizing nucleosides. Semin. Radiat. Oncol., 11: 270-280, 2001.[Medline]
  101. Haydon R. C., Zhou L., Feng T., Breyer B., Cheng H., Jiang W., Ishikawa A., Peabody T., Montag A., Simon M. A., He T. C. Nuclear receptor agonists as potential differentiation therapy agents for human osteosarcoma. Clin. Cancer Res., 8: 1288-1294, 2002.[Abstract/Free Full Text]
  102. Modak S., Kramer K., Gultekin S. H., Guo H. F., Cheung N. K. Monoclonal antibody 8H9 targets a novel cell surface antigen expressed by a wide spectrum of human solid tumors. Cancer Res., 61: 4048-4054, 2001.[Abstract/Free Full Text]
  103. Cheung N. K., Modak S., Vickers A., Knuckles B. Orally administered ß-glucans enhance anti-tumor effects of monoclonal antibodies. Cancer Immunol. Immunother., 51: 557-564, 2002.[Medline]
  104. Krause A., Guo H. F., Latouche J. B., Tan C., Cheung N. K., Sadelain M. Antigen-dependent CD28 signaling selectively enhances survival and proliferation in genetically modified activated human primary T lymphocytes. J. Exp. Med., 188: 619-626, 1998.[Abstract/Free Full Text]
  105. Matsubara S., Wada Y., Gardner T. A., Egawa M., Park M. S., Hsieh C. L., Zhau H. E., Kao C., Kamidono S., Gillenwater J. Y., Chung L. W. A conditional replication-competent adenoviral vector. Ad-OC-E1a, to cotarget prostate cancer and bone stroma in an experimental model of androgen-independent prostate cancer bone metastasis. Cancer Res., 61: 6012-6019, 2001.[Abstract/Free Full Text]
  106. Shirakawa T., Ko S. C., Gardner T. A., Cheon J., Miyamoto T., Gotoh A., Chung L. W., Kao C. In vivo suppression of osteosarcoma pulmonary metastasis with intravenous osteocalcin promoter-based toxic gene therapy. Cancer Gene Ther., 5: 274-280, 1998.[Medline]
  107. Cheon J., Ko S. C., Gardner T. A., Shirakawa T., Gotoh A., Kao C., Chung L. W. Chemogene therapy: osteocalcin promoter-based suicide gene therapy in combination with methotrexate in a murine osteosarcoma model. Cancer Gene Ther., 4: 359-365, 1997.[Medline]
  108. Kubo H., Gardner T. A., Wada Y., Koeneman K. S., Gotoh A., Yang L., Kao C., Lim S. D., Amin M. B., Yang H., Black M. E., Matsubara S., Nakagawa M., Gillenwater J. Y., Zhau H. E., Chung L. W. Phase I dose escalation clinical trial of adenovirus vector carrying osteocalcin promoter-driven herpes simplex virus thymidine kinase in localized and metastatic hormone-refractory prostate cancer. Hum. Gene Ther., 14: 227-241, 2003.[Medline]
  109. Anderson P. M., Markovic S. N., Sloan J. A., Clawson M. L., Wylam M., Arndt C. A., Smithson W. A., Burch P., Gornet M., Rahman E. Aerosol granulocyte macrophage-colony stimulating factor: a low toxicity, lung-specific biological therapy in patients with lung metastases. Clin. Cancer Res., 5: 2316-2323, 1999.[Abstract/Free Full Text]
  110. Withrow S. J., Straw R. C. Effects of cis-diamminedichloroplatinum II released from D. L-polylactic acid implanted adjacent to cortical allografts in dogs. Vet. Surg., 23: 341-346, 1994.[Medline]
  111. Dernell W. S., Withrow S. J., Straw R. C., Powers B. E., Drekke J. H., Lafferty M. Intracavitary treatment of soft tissue sarcomas in dogs using cisplatin in a biodegradable polymer. Anticancer Res., 17: 4499-505, 1997.[Medline]
  112. Lana S. E., Dernell W. S., LaRue S. M., Lafferty M. J., Douple E. B., Brekke J. H., Withrow S. J. Slow release cisplatin combined with radiation for the treatment of canine nasal tumors. Vet. Radiol. Ultrasound, 38: 474-478, 1997.[Medline]
  113. Allgrove J. Use of bisphosphonates in children and adolescents. J. Pediatric Endocrinol. Metab., 15 (Suppl. 3): 921-928, 2002.
  114. Giuliani N., Pedrazzoni M., Passeri G., Girasole G. Bisphosphonates inhibit IL-6 production by human osteoblast-like cells. Scand. J. Rheumatol., 27: 38-41, 1998.[Medline]
  115. Diel I. J., Meyberg G. C., Gollan C., Bode S., Wallwiener D., Bastert G. Reduction in new metastases in breast cancer with adjuvant clodronate treatment. Breast Cancer Res. Treat., 49: 145-154, 1998.[Medline]
  116. Diel I. J., Solomayer E. F., Bastert G. Bisphosphonates and the prevention of metastasis: first evidences from preclinical and clinical studies. Cancer (Phila.), 88: 3080-3088, 2000.
  117. Diel I. J. Bisphosphonates in the prevention of bone metastases: current evidence. Semin. Oncol., 28: 75-80, 2001.[Medline]
  118. Ecklund K., Laor T., Goorin A. M., Connolly L. P., Jaramillo D. Methotrexate osteopathy in patients with osteosarcoma. Radiology, 202: 543-547, 1997.[Abstract]
  119. Healey J. The epidemic of chemotherapy-related osteoporosis. Curr. Opin. Orthop., 10: 331-333, 1999.
  120. Pfeilschifter J., Diel I. J. Osteoporosis due to cancer treatment: pathogenesis and management. J. Clin. Oncol., 18: 1570-1593, 2000.[Abstract/Free Full Text]
  121. Brumsen C., Hamdy N. A., Papapoulos S. E. Long-term effects of bisphosphonates on the growing skeleton. Studies of young patients with severe osteoporosis. Medicine, 76: 266-283, 1997.[Medline]
  122. Kerbel R., Viloria-Petit A., Klement G., Rak J. ‘Accidental’ anti-angiogenic drugs: anti-oncogene directed signal transduction inhibitors and conventional chemotherapeutic agents as examples. Eur. J. Cancer, 36: 1248-1257, 2000.[Medline]
  123. Gately S. The contributions of cyclooxygenase-2 to tumor angiogenesis. Cancer Metastasis Rev., 19: 19-27, 2000.[Medline]
  124. Stempak D., Gammon J., Klein J., Moriarty P., Halton J., Koren G., Baruchel S. Safety and pharmacokinetics of celecoxib when combined with metronomic administration of low dose chemotherapy for anti-angiogenic therapy of pediatric recurrent solid tumors. Proc. Am. Assoc. Cancer Res., 42: 19A 2001.
  125. Stempak D., Gammon J., Klein J., Koren G., Baruchel S. Single-dose and steady-state pharmacokinetics of celecoxib in children. Clin. Pharmacol. Ther., 72: 490-497, 2002.[Medline]

Back to Scientific Conferences Home


Last Reviewed: May 15, 2003
Back to Top
Back to Top