throbber
An Integrated Genomic Analysis of Human Glioblastoma Multiforme
`D. Williams Parsons,1,2* Siân Jones,1* Xiaosong Zhang,1* Jimmy Cheng-Ho Lin,1* Rebecca J. Leary,1*
`Philipp Angenendt,1* Parminder Mankoo,3 Hannah Carter,3 I-Mei Siu,4 Gary L. Gallia,4 Alessandro Olivi,4
`Roger McLendon,5 B. Ahmed Rasheed,5 Stephen Keir,5 Tatiana Nikolskaya,6 Yuri Nikolsky,7 Dana A.
`Busam,8 Hanna Tekleab,8 Luis A. Diaz Jr.,1 James Hartigan,9 Doug R. Smith,9 Robert L. Strausberg,8 Suely
`Kazue Nagahashi Marie,10 Sueli Mieko Oba Shinjo,10 Hai Yan,5 Gregory J. Riggins,4 Darell D. Bigner,5
`Rachel Karchin,3 Nick Papadopoulos,1 Giovanni Parmigiani,1 Bert Vogelstein,1† Victor E. Velculescu,1†
`Kenneth W. Kinzler1†
`1Ludwig Center for Cancer Genetics and Therapeutics, and The Howard Hughes Medical Institute at Johns Hopkins Kimmel
`Cancer Center, Baltimore, MD 21231, USA. 2Department of Pediatrics, Section of Hematology-Oncology, Baylor College of
`Medicine, Houston TX 77030, USA. 3Department of Biomedical Engineering, Institute of Computational Medicine, Johns
`Hopkins Medical Institutions, Baltimore, MD 21218, USA. 4Department of Neurosurgery, Johns Hopkins Medical Institutions,
`Baltimore, MD 21231, USA. 5Department of Pathology, Pediatric Brain Tumor Foundation, and Preston Robert Tisch Brain
`Tumor Center at Duke University Medical Center, Durham, NC 27710, USA. 6Vavilov Institute for General Genetics, Moscow
`B333, 117809, Russia. 7GeneGo, Inc., St. Joseph, MI 49085, USA. 8J. Craig Venter Institute, Rockville, MD 20850, USA.
`9Agencourt Bioscience Corporation, Beverly, MA 01915, USA. 10Department of Neurology, School of Medicine, University of
`São Paulo, São Paulo, Brazil.
`
`*These authors contributed equally to this work.
`
`†To whom correspondence should be addressed. E-mail: bertvog@gmail.com (B.V.); velculescu@jhmi.edu (V.E.V.);
`kinzlke@welch.jhu.edu (K.W.K.)
`
`Glioblastoma multiforme (GBM) is the most common and
`lethal type of brain cancer. To identify the genetic
`alterations in GBMs, we sequenced 20,661 protein coding
`genes, determined the presence of amplifications and
`deletions using high-density oligonucleotide arrays, and
`performed gene expression analyses using next-generation
`sequencing technologies in 22 human tumor samples. This
`comprehensive analysis led to the discovery of a variety of
`genes that were not known to be altered in GBMs. Most
`notably, we found recurrent mutations in the active site of
`isocitrate dehydrogenase 1 (IDH1) in 12% of GBM
`patients. Mutations in IDH1 occurred in a large fraction
`of young patients and in most patients with secondary
`GBMs, and were associated with an increase in overall
`survival. These studies demonstrate the value of unbiased
`genomic analyses in the characterization of human brain
`cancer and identify a potentially useful genetic alteration
`for the classification and targeted therapy of GBMs.
`Malignant gliomas are the most frequent and lethal cancers
`originating in the central nervous system. The most
`biologically aggressive subtype is glioblastoma multiforme
`(GBM; WHO grade IV astrocytoma), a tumor associated with
`a dismal prognosis (1). The current standard of care for GBM
`patients—surgical resection followed by adjuvant radiation
`
`therapy and chemotherapy with the oral alkylating agent
`temozolomide—produces a median survival of only 15
`months (2). Historically, GBMs have been categorized into
`two groups (“primary” and “secondary”) on the basis of
`clinical presentation (3). Secondary GBMs are defined as
`cancers that have clinical, radiologic, or histopathologic
`evidence of malignant progression from a preexisting lower-
`grade tumor, while primary GBMs have no such history and
`present at diagnosis as advanced cancers (4). Clinical
`differences have been reported between the two groups, with
`secondary GBMs occurring less frequently (~5% of GBMs)
`and predominantly in younger patients (median age ~45 years
`vs. ~60 years for primary GBM) (5, 6). The histopathologic
`findings of primary and secondary GBMs are
`indistinguishable, and the prognosis does not appear to be
`significantly different after adjustment for age (5, 6).
`Substantial research effort has focused on the
`identification of genetic alterations in GBMs that might help
`define subclasses of GBM patients with differing prognoses
`and/or response to specific therapies (7). Distinctions between
`the genetic lesions found in primary and secondary GBMs
`have been made, with TP53 mutations occurring more
`commonly in secondary GBMs and EGFR amplifications and
`PTEN mutations occurring more frequently in primary GBMs
`
`/ www.sciencexpress.org / 4 September 2008 / Page 1 / 10.1126/science.1164382
`
`Rigel Exhibit 1015
`Page 1 of 13
`
`

`

`(6, 8, 9); however, none of these alterations are sufficiently
`specific to distinguish between primary and secondary
`GBMs. This issue is further confounded by the possibility
`that a fraction of GBMs designated as primary tumors may
`follow a sequence of genetic events similar to that of
`secondary lesions but not come to clinical attention until
`malignant progression to a GBM has occurred.
`The comprehensive elucidation of genetic alterations in
`GBMs could provide novel targets that might be used for
`diagnostic, prognostic, or therapeutic purposes as well as
`identify subgroups of patients that preferentially respond to
`particular targeted therapies. The determination of the human
`genome sequence and improvements in sequencing and
`bioinformatic technologies have recently permitted genome-
`wide sequence analyses in human cancers. We have
`previously studied the genomes of 11 breast and 11 colorectal
`cancers by determining the sequence of the more than 18,000
`Consensus Coding Sequence (CCDS) and Reference
`Sequence (RefSeq) genes (10, 11). Here, we have analyzed
`20,661 protein coding genes in 22 human GBM samples. To
`complement these sequencing data, we have also performed a
`genome-wide analysis of focal copy number alterations,
`including amplifications and homozygous deletions, using
`high-density oligonucleotide microarrays on the same GBM
`tumors. Finally, we have examined the expression profiles of
`these same samples using serial analysis of gene expression
`(SAGE) and next-generation sequencing technologies.
`
`Sequencing strategy. We extended our previous
`sequencing strategy for identification of somatic mutations to
`include 23,219 transcripts from 20,661 genes (fig. S1). These
`included 2783 additional genes from the Ensembl databases
`that were not present in the CCDS or RefSeq databases
`analyzed in the previous studies (10, 11). In addition, we
`redesigned PCR primers for regions of the genome that (i)
`were difficult to PCR amplify in prior studies; or (ii) were
`found to share significant identity with other human or mouse
`sequences. The combination of these new, redesigned, and
`existing primers sequences resulted in a total of 208,311
`primer pairs (table S1) that were successfully used for
`sequence analysis of the coding exons of these genes.
`Twenty-two GBM samples (table S2) were selected for
`PCR sequence analysis, consisting of 7 samples extracted
`directly from patient tumors and 15 samples passaged in nude
`mice as xenografts. In the first stage of this analysis, called
`the Discovery Screen, the primer pairs were used to amplify
`and sequence 175,471 coding exons and adjacent intronic
`splice donor and acceptor sequences in 22 GBM samples and
`one matched normal sample. The data were assembled for
`each amplified region and evaluated using stringent quality
`criteria (12), resulting in successful amplification and
`sequencing of 95.0% of targeted amplicons and 93.0% of
`
`targeted bases in the 22 tumors (Table 1). A total of 689 Mb
`of sequence data was generated in this fashion. The amplicon
`traces were analyzed using automated approaches to identify
`changes in the tumor sequences that were not present in the
`reference sequences of each gene. Alterations present in the
`normal control sample and in single nucleotide polymorphism
`(SNP) databases were then removed from further analyses.
`The remaining sequence traces of potential alterations were
`visually inspected to remove false-positive mutation calls
`generated by the automated software. All exons containing
`putative mutations were then reamplified and sequenced in
`both the affected tumor and the matched normal DNA
`sample. This process allowed us to confirm the presence of
`the mutation in the tumor sample and determine whether the
`alteration was somatic (i.e. tumor-specific) or was present in
`the germline. All putative somatic mutations were examined
`computationally and experimentally to confirm that the
`alterations did not arise through the aberrant coamplification
`of related gene sequences (12).
`
`Analysis of sequence alterations. Analysis of the
`identified somatic mutations revealed that one tumor (Br27P),
`from a patient previously treated with radiation therapy and
`temozolomide, had 17-fold more alterations than any of the
`other 21 patients (table S3). The mutation spectrum of this
`sample was also dramatically different from those of the other
`GBM patients (12) and was consistent with previous
`observations of a hypermutation phenotype in glioma samples
`of patients treated with temozolomide (13, 14). After
`removing Br27P from consideration, we found that 685 genes
`(3.3% of the 20,661 genes analyzed) contained at least one
`nonsilent somatic mutation. The vast majority of these
`alterations were single-base substitutions (94%), while the
`others were small insertions, deletions, or duplications (Table
`1). The 993 somatic mutations were observed to be
`distributed relatively evenly among the 21 remaining tumors
`(table S3), with a mean of 47 mutations per tumor,
`representing 1.51 mutations per Mb of GBM tumor genome
`sequenced. The six DNA samples extracted directly from
`patient tumors had smaller numbers of mutations than those
`obtained from xenografts, likely because of the masking
`effect of nonneoplastic cells in the former. It has previously
`been shown that cell lines and xenografts provide the optimal
`template DNA for cancer genome sequencing analyses (15)
`and that they faithfully represent the alterations present in the
`original tumors (16). Both the total number and frequency of
`sequence alterations in GBMs were substantially smaller than
`the number and frequency of such alterations observed in
`colorectal or breast cancers, and slightly less than in
`pancreatic cancers (10, 11, 17). The most likely explanation
`for this difference is the reduced number of cell generations
`in glial cells prior to the onset of neoplasia (18).
`
`/ www.sciencexpress.org / 4 September 2008 / Page 2 / 10.1126/science.1164382
`
`Rigel Exhibit 1015
`Page 2 of 13
`
`

`

`We further evaluated a set of 21 mutated genes identified
`in the Discovery Screen in a second screen, called a
`Prevalence Screen, comprising an additional 83 GBMs with
`well-documented clinical histories (table S2). The 21 genes
`selected were mutated in at least two Discovery Screen
`tumors and had mutation frequencies of >10 mutations per
`Mb of tumor DNA sequenced. Nonsilent somatic mutations
`were identified in 16 of these 21 genes in the additional tumor
`samples (table S4). The mutation frequency of all analyzed
`genes in the Prevalence Screen was 24 mutations per Mb of
`tumor DNA, markedly increased from the overall mutation
`frequency in the Discovery Screen of 1.5 mutations per Mb
`(P < 0.001, binomial test). Additionally, the observed ratio of
`nonsilent to silent mutations among mutations in the
`Prevalence Screen was 14.8:1, substantially higher than the
`3.1:1 ratio that was observed in the Discovery Screen (P <
`0.001, binomial test). The increased mutation frequency and
`higher fraction of nonsilent mutations suggested that genes
`mutated in the Prevalence Screen were enriched for genes
`that actively contributed to tumorigenesis.
`In addition to the frequency of mutations in a gene, the
`type of mutation can provide information useful for
`evaluating its potential role in disease (19). The likely effect
`of missense mutations can be assessed through evaluation of
`the mutated residue by evolutionary or structural means. To
`evaluate missense mutations, we developed an algorithm (LS-
`MUT) that employs machine learning of 58 predictive
`features based on evolutionary conservation and the physical-
`chemical properties of amino acids involved in the alteration
`(12). Approximately 15% of the missense mutations
`evaluated were predicted to have a statistically significant
`effect on protein function when assessed by this method
`(table S3). We also were able to make structural models of
`244 of the 870 missense mutations identified in this study
`(20). In each case, the model was based on x-ray
`crystallography or nuclear magnetic resonance spectroscopy
`of the normal protein or a closely related homolog. This
`analysis showed that 35 of the missense mutations are located
`close to a domain interface or substrate-binding site and thus
`are likely to impact protein function [(links to structural
`models available in (12)].
`
`Analysis of copy number changes. The same tumors
`were then evaluated for copy number alterations through
`genomic hybridization of DNA samples to Illumina SNP
`arrays containing ~1 million probes (21). We have recently
`developed a sensitive and specific approach for the
`identification of focal amplifications resulting in 12 or more
`copies per nucleus (6-fold or greater amplification compared
`to the diploid genome) as well as deletions of both copies of a
`gene (homozygous deletions) using such arrays (22). Unlike
`larger chromosomal aberrations, such focused alterations can
`
`be used to identify underlying candidate genes in these
`regions.
`We identified a total of 147 amplifications (table S5) and
`134 homozygous deletions (table S6) in the 22 samples used
`in the Discovery Screen (Table 1). Although the number of
`amplifications was similar in samples extracted from patient
`tumors and those that had been passaged as xenografts, the
`latter samples allowed detection of a larger number of
`homozygous deletions (average of 8.0 deletions per sample in
`the xenografts versus 2.2 per sample in the tumors). These
`observations are consistent with previous reports
`documenting the difficulty of identifying homozygous
`deletions in samples containing contaminating normal DNA
`(23) and highlight the importance of using purified human
`tumor cells, such as those present in xenografts or cell lines,
`for genomic analyses.
`
`Integration of sequencing, copy number and expression
`analyses. Mutations that arise during tumorigenesis may
`provide a selective advantage to the tumor cell (driver
`mutations) or have no net effect on tumor growth (passenger
`mutations). The mutational data obtained from sequencing
`and analysis of copy number alterations were integrated in
`order to identify GBM candidate cancer genes (CAN-genes)
`that are most likely to be drivers and therefore worthy of
`further investigation. To determine if a gene was likely to
`harbor driver mutations, we compared the number and type of
`mutations observed (including sequence changes,
`amplifications and homozygous deletions) and determined the
`probability that these alterations would result from passenger
`mutation rates alone (12) (fig. S1).
`The CAN-genes, together with their passenger
`probabilities, are listed in table S7. The CAN-genes included
`several with well-established roles in gliomas, including
`TP53, PTEN, CDKN2A, RB1, EGFR, NF1, PIK3CA and
`PIK3R1 (24–34). Of these genes, the most frequently altered
`were CDKN2A (altered in 50% of GBMs), TP53, EGFR, and
`PTEN (altered in 30 to 40%), NF1, CDK4 and RB1 (altered in
`12 to 15%), and PIK3CA and PIK3R1 (altered in 8 to 10%)
`(Table 2). Overall, these frequencies, which are similar to or
`in some cases higher than those previously reported, validate
`the sensitivity of our approach for detecting somatic
`alterations.
`Through analysis of additional gene members within cell
`signaling pathways affected by these genes, we identified
`alterations of critical genes in the TP53 pathway (TP53,
`MDM2, MDM4), the RB1 pathway (RB1, CDK4, CDKN2A),
`and the PI3K/PTEN pathway (PIK3CA, PIK3R1, PTEN,
`IRS1). These alterations affected pathways in a majority of
`tumors (64%, 68%, and 50%, respectively) and in all cases
`but one, mutations within each tumor affected only a single
`
`/ www.sciencexpress.org / 4 September 2008 / Page 3 / 10.1126/science.1164382
`
`Rigel Exhibit 1015
`Page 3 of 13
`
`

`

`member of each pathway in a mutually exclusive manner (P <
`0.05) (Table 3).
`Systematic analyses of functional gene groups and
`pathways contained within the well-annotated MetaCore
`database (35) identified enrichment of alterations in a variety
`of cellular processes in GBMs, including additional members
`of the TP53 and PI3K/PTEN pathways. Many of the
`pathways identified were similar to core signaling pathways
`found to be altered in pancreas, colorectal, and breast tumors,
`such as those regulating control of cellular growth, apoptosis
`and cell adhesion (17, 22, 36). However, several pathways
`were enriched only in GBMs. These included channels
`involved in transport of sodium, potassium and calcium ions
`as well as nervous system-specific cellular pathways such as
`synaptic transmission, transmission of nerve impulses, and
`axonal guidance (table S8). Mutations in these latter
`pathways may represent a subversion of normal glial cell
`processes to promote dysregulated growth and invasion.
`Gene expression patterns can inform the analysis of
`pathways because they can reflect epigenetic alterations not
`detectable by sequencing or copy number analyses. To
`analyze the transcriptome of GBMs, we performed SAGE
`(serial analysis of gene expression) (37, 38) on all GBM
`samples for which RNA was available (total of 18 samples)
`as well as on two independent normal brain RNA controls.
`When combined with sequencing-by-synthesis methods (39–
`42), SAGE provides a highly quantitative and sensitive
`measure of gene expression. We first used the transcript
`analysis to help identify previously uncharacterized target
`genes from the amplified and deleted regions that were
`revealed by our study. In tables S5 and S6, a candidate target
`gene could be identified within several of these regions
`through the use of the mutational as well as transcriptional
`data. Second, we used the transcript analysis to help identify
`genes that were differentially expressed in GBMs compared
`to normal brain. A large number of genes (143) were
`expressed on average at 10-fold higher levels in the GBMs.
`Among the overexpressed genes, 16 encoded proteins that are
`predicted to be secreted or expressed on the cell surface,
`suggesting new opportunities for diagnostic and therapeutic
`applications. Finally, we assessed whether the gene sets
`implicated in the pathways enriched for genetic alterations
`were also altered through expression changes. Notably, the
`gene sets in these pathways were more highly enriched for
`differentially expressed genes than the remaining sets (P <
`0.001) (12). These expression data thus independently
`highlight the potential importance of these pathways in the
`development of GBMs.
`
`High frequency alterations of IDH1 in GBM. The CAN-
`gene list (table S7) included a number of individual genes that
`had not previously been linked to GBMs. The most frequently
`
`mutated of these genes, IDH1 on chromosome 2q33, encodes
`isocitrate dehydrogenase 1, which catalyzes the oxidative
`carboxylation of isocitrate to (cid:68)-ketoglutarate, resulting in the
`production of NADPH. Of the five isocitrate dehydrogenase
`proteins encoded in the human genome, four are localized to
`the mitochondria while only IDH1 is localized within the
`cytoplasm and peroxisomes (43). The IDH1 protein forms an
`asymmetric homodimer (44), and is thought to play a
`substantial role in cellular control of oxidative damage
`through generation of NADPH (45, 46). None of the other
`IDH genes were found to be genetically altered in our
`analysis.
`IDH1 was somatically mutated in five of the 22 GBM
`tumors in the Discovery Screen. Surprisingly, all five had the
`same heterozygous point mutation, a change of a guanine to
`an adenine at position 395 of the IDH1 transcript (G395A),
`leading to the replacement of an arginine with a histidine at
`amino acid residue 132 of the protein (R132H). In our prior
`study of colorectal cancers, this same codon was mutated in a
`single case through alteration of the adjacent nucleotide,
`resulting in a R132C amino acid change (10). Five GBMs
`evaluated in our Prevalence Screen were found to have
`heterozygous somatic R132H mutations and an additional
`two tumors had a third distinct somatic mutation affecting the
`same amino acid residue, R132S (fig. S2 and Table 4). In
`addition to the Discovery and Prevalence Screen samples, 44
`other GBMs were analyzed for IDH1 mutations , revealing
`six tumors with somatic mutations affecting R132. In total, 18
`of 149 GBMs (12%) analyzed had alterations in IDH1. The
`R132 residue is conserved in all known species and is
`localized to the substrate binding site, where it forms
`hydrophilic interactions with the alpha-carboxylate of
`isocitrate (Fig. 1) (44, 47).
`Several important observations were made about IDH1
`mutations and their potential clinical significance. First,
`mutations in IDH1 preferentially occurred in younger GBM
`patients, with a mean age of 33 years for IDH1-mutated
`patients, as opposed to 53 years for patients with wild-type
`IDH1 (P < 0.001, t test, Table 4). In patients under 35 years
`of age, nearly 50% (9 of 19) had mutations in IDH1. Second,
`mutations in IDH1 were found in nearly all of the patients
`with secondary GBMs (mutations in 5 of 6 secondary GBM
`patients, as compared to 7 of 99 patients with primary GBMs,
`P < 0.001, binomial test). Third, patients with IDH1
`mutations had a significantly improved prognosis, with a
`median overall survival of 3.8 years as compared to 1.1 years
`for patients with wild-type IDH1 (Fig. 2, P < 0.001, log-rank
`test). Although both younger age and mutated TP53 are
`known to be positive prognostic factors for GBM patients,
`this association between IDH1 mutation and improved
`survival was noted even in the subgroup of young patients
`with TP53 mutations (P < 0.02, log-rank test).
`
`/ www.sciencexpress.org / 4 September 2008 / Page 4 / 10.1126/science.1164382
`
`Rigel Exhibit 1015
`Page 4 of 13
`
`

`

`and the fact that all mutations observed to date were
`heterozygous (without any evidence of loss of the second
`allele through LOH). Interestingly, enzymatic studies have
`shown that in vitro engineered substitution of arginine at
`residue 132 with a different amino acid than observed in
`patients (glutamate) results in a catalytically inactive enzyme,
`suggesting a critical role for this residue (48). Further
`biochemical and molecular analyses will be needed to
`determine the effect of alterations of IDH1 on enzymatic
`activity and cellular phenotype.
`Regardless of the specific molecular consequences of
`IDH1 alterations, detection of mutations in IDH1 is likely to
`be clinically useful. Although significant effort has focused
`on the identification of characteristic genetic lesions in
`primary and secondary GBMs, the altered genes identified to
`date are not optimal for this purpose (5). Our study revealed
`IDH1 mutation to be a novel and potentially more specific
`marker for secondary GBM. One hypothesis is that IDH1
`alterations identify a biologically specific subgroup of GBM
`patients, including both patients who would be classified as
`having secondary GBMs as well as a subpopulation of
`primary GBM patients with a similar tumor biology and more
`protracted clinical course (Table 4). Interestingly, patients
`with IDH1 mutations had a very high frequency of TP53
`mutation and a very low frequency of mutations in other
`commonly altered GBM genes (Table 4). Patients with
`mutated IDH1 also had distinct clinical characteristics,
`including younger age and a significantly improved clinical
`prognosis (Table 4). It is conceivable that new treatments
`could be designed to take advantage of IDH1 alterations in
`these patients, as inhibition of a different IDH enzyme
`(mitochondrial IDH2) has recently been shown to result in
`increased sensitivity of tumor cells to a variety of
`chemotherapeutic agents (49). In summary, the discovery of
`IDH1 and other genes previously not known to play a role in
`human tumors (table S7) validates the utility of genome-wide
`genetic analysis of tumors in general and opens new avenues
`of basic and clinical brain tumor research.
`
`Discussion. The data resulting from this integrated
`analysis of mutations and copy number alterations have
`provided a novel view of the genetic landscape of
`glioblastomas. Like all large-scale genetic analyses, our study
`has limitations. We did not assess certain molecular
`alterations, including chromosomal translocations and
`epigenetic changes. However, our large scale expression
`studies should have identified any genes that were
`differentially expressed through these mechanisms (table S9).
`Additionally, we focused on copy number changes that were
`focal amplifications or homozygous deletions, as these have
`historically been most useful in identifying cancer genes. The
`array data we have generated can also be analyzed to
`determine loss of heterozygosity (LOH) or low-amplitude
`regions of copy number gains, but such changes cannot
`generally be used to pinpoint new candidate cancer genes.
`Finally, the samples directly extracted from patient tumors
`contained small amounts of contaminating normal tissue,
`which limited our ability to detect homozygous deletions and
`to a lesser extent, somatic mutations, in those specific tumors.
`Despite these limitations, our study provides a number of
`important genetic and clinical insights into GBMs. First, it
`revealed that some of the pathways known to be altered in
`GBMs affect a larger fraction of genes and patients than
`previously anticipated. A majority of the tumors analyzed had
`alterations in genes encoding components of each of the
`TP53, RB1, and PI3K pathways. The fact that all but one of
`the cancers with mutations in members of a pathway did not
`have alterations in other members of the same pathway
`suggests that such alterations are functionally equivalent in
`tumorigenesis. Second, these results have identified a variety
`of new genes and signaling pathways not previously
`implicated in GBMs (table S7 and S8). Some of these
`pathways were found to be altered in previous genome-wide
`analyses of pancreatic, breast and colorectal cancers and may
`represent core processes that underlie human tumorigenesis
`(17, 22, 36). A number of the signaling pathways mutated or
`altered through expression differences in GBMs appear to be
`involved in nervous system signaling processes and represent
`novel and potentially useful aspects of GBM biology.
`The comprehensive nature of our study allowed us to
`identify IDH1 as an unexpected target of genetic alteration in
`patients with GBM. All mutations in this gene resulted in
`amino acid substitutions at position 132, an evolutionarily
`conserved residue located within the isocitrate binding site
`(44). The recurrent nature of the mutations is reminiscent of
`activating alterations in oncogenes such as BRAF, KRAS, and
`PIK3CA. Our speculation that this sequence change is an
`activating mutation is strengthened by the absence of
`inactivating changes (e.g., frameshift or stop mutations), the
`absence of other alterations in key residues of the active site,
`
`References and Notes
`1. D. N. Louis et al., Acta Neuropathol 114, 97 (2007).
`2. R. Stupp et al., N Engl J Med 352, 987 (2005).
`3. H. Scherer, American Journal of Cancer 40, 159 (1940).
`4. P. Kleihues, H. Ohgaki, Neuro Oncol 1, 44 (1999).
`5. H. Ohgaki, P. Kleihues, Am J Pathol 170, 1445 (2007).
`6. H. Ohgaki et al., Cancer Res 64, 6892 (2004).
`7. I. K. Mellinghoff et al., N Engl J Med 353, 2012 (2005).
`8. E. A. Maher et al., Cancer Res 66, 11502 (2006).
`9. C. L. Tso et al., Cancer Res 66, 159 (2006).
`10. T. Sjöblom et al., Science 314, 268 (2006).
`11. L. D. Wood et al., Science 318, 1108 (2007).
`12. See Supporting Online Material.
`13. D. P. Cahill et al., Clin Cancer Res 13, 2038 (2007).
`/ www.sciencexpress.org / 4 September 2008 / Page 5 / 10.1126/science.1164382
`
`Rigel Exhibit 1015
`Page 5 of 13
`
`

`

`14. C. Hunter et al., Cancer Res 66, 3987 (2006).
`15. J. M. Winter, J. R. Brody, S. E. Kern, Cancer Biol Ther 5,
`360 (2006).
`16. S. Jones et al., Proc Natl Acad Sci U S A 105, 4283
`(2008).
`17. S. Jones et al., Science, this issue (2008).
`18. R. Kraus-Ruppert, J. Laissue, H. Burki, N. Odartchenko, J
`Comp Neurol 148, 211 (1973).
`19. P. C. Ng, S. Henikoff, Nucleic Acids Res 31, 3812 (2003).
`20. R. Karchin (2008). Structural models of mutants
`identified in glioblastomas.
`http://karchinlab.org/Mutants/CAN-
`genes/brain/GBM.html
`21. F. J. Steemers et al., Nat Methods 3, 31 (2006).
`22. R. J. Leary et al., Proc Natl Acad Sci U S A, in press.
`23. P. Cairns et al., Nat Genet 11, 210 (1995).
`24. J. M. Nigro et al., Nature 342, 705 (1989).
`25. J. Li et al., Science 275, 1943 (1997).
`26. K. Ueki et al., Cancer Res 56, 150 (1996).
`27. A. J. Wong et al., Proc Natl Acad Sci U S A 84, 6899
`(1987).
`28. A. J. Wong et al., Proc Natl Acad Sci U S A 89, 2965
`(1992).
`29. L. Frederick, X. Y. Wang, G. Eley, C. D. James, Cancer
`Res 60, 1383 (2000).
`30. Y. Li et al., Cell 69, 275 (1992).
`31. G. Thiel et al., Anticancer Res 15, 2495 (1995).
`32. Y. Samuels et al., Science 304, 554 (2004).
`33. D. K. Broderick et al., Cancer Res 64, 5048 (2004).
`34. G. L. Gallia et al., Mol Cancer Res 4, 709 (2006).
`35. S. Ekins, Y. Nikolsky, A. Bugrim, E. Kirillov, T.
`Nikolskaya, Methods Mol Biol 356, 319 (2007).
`36. J. Lin et al., Genome Res 17, 1304 (2007).
`37. V. E. Velculescu, L. Zhang, B. Vogelstein, K. W. Kinzler,
`Science 270, 484 (1995).
`38. S. Saha et al., Nat Biotechnol 20, 508 (2002).
`39. M. Sultan et al., Science (2008).
`40. R. Lister et al., Cell 133, 523 (2008).
`41. A. Mortazavi, B. A. Williams, K. McCue, L. Schaeffer,
`B. Wold, Nat Methods 5, 621 (2008).
`42. R. Morin et al., Biotechniques 45, 81 (2008).
`43. B. V. Geisbrecht, S. J. Gould, J Biol Chem 274, 30527
`(1999).
`44. X. Xu et al., J Biol Chem 279, 33946 (2004).
`45. S. M. Lee et al., Free Radic Biol Med 32, 1185 (2002).
`46. S. Y. Kim et al., Mol Cell Biochem 302, 27 (2007).
`47. A. Nekrutenko, D. M. Hillis, J. C. Patton, R. D. Bradley,
`R. J. Baker, Mol Biol Evol 15, 1674 (1998).
`48. G. T. Jennings, K. I. Minard, L. McAlister-Henn,
`Biochemistry 36, 13743 (1997).
`49. I. S. Kil, S. Y. Kim, S. J. Lee, J. W. Park, Free Radic Biol
`Med 43, 1197 (2007).
`
`50. E. F. Pettersen et al., J Comput Chem 25, 1605 (2004).
`51. We thank N. Silliman, J. Ptak, L. Dobbyn, and M.
`Whalen for assistance with PCR amplification; D. Lister,
`L. J. Ehinger, D. L. Satterfield, J. D. Funkhouser, and P.
`Killela for assistance with DNA purification; T. Sjöblom
`with for assistance with database management; the
`Agencourt sequencing team for assistance with automated
`sequencing; and C.-S. Liu and the SoftGenetics team for
`their assistance with mutation detection analyses. This
`project was carried out under the auspices of the Ludwig
`Brain Tumor Initiative and was supported by The Virginia
`and D. K. Ludwig Fund for Cancer Research, NIH grants
`CA121113, NS052507, CA43460, CA 57345, CA62924,
`CA09547, 5P50-NS-20023, CA108786, CA11898, The
`Pew Charitable Trusts, The Pediatric Brain Tumor
`Foundation Institute, The Hirschhorn Foundation, Alex’s
`Lemonade Stand Foundation, The American Brain Tumor
`Association, The American Society of Clinical Oncology,
`The Brain Tumor Research Fund and Beckman Coulter
`Corporation. Under separate licensing agreements between
`the Johns Hopkins University and Genzyme, Beckman
`Coulter, and Exact Sciences Corporations, B.V., V.E.V.,
`and K.W.K. are entitled to a share of royalties received by
`the University on sales of products related to research
`described in this paper. These authors and the University
`own Genzyme and Exact Sciences stock, which is subject
`to certain restrictions under University policy. The terms
`of these arrangements are managed by the Johns Hopkins
`University in accordance with its conflict-of-interest
`policies.
`
`Supporting Online Material
`www.sciencemag.org/cgi/content/full/1164382/DC1
`Materials and Methods
`Figs. S1 and S2
`Tables S1 to S9
`References
`
`7 August 2008; accepted 27 August 2008
`Published online 4 September 2008;
`10.1126/science.1164382
`Include this information when citing this paper.
`
`Fig. 1. Structure of the active site of IDH1. The crystal
`structure of the human cytosolic NADP(+)–dependent IDH is
`shown in ribbon format (PDBID: 1T0L) (44). The active cleft
`of IDH1 consists of a NADP-binding site and the isocitrate-
`meta

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket