throbber
Letter
`
`pubs.acs.org/acsmedchemlett
`
`†
`
`∥
`
`‡
`

`
`Discovery of the First Potent Inhibitors of Mutant IDH1 That Lower
`Tumor 2‑HG in Vivo
`Janeta Popovici-Muller,*,†
`Shunqi Yan,
`Jeremy M. Travins,
`Francesco G. Salituro,
`Jeffrey O. Saunders,
`†
`†
`†
`†
`†
`⊥
`Fang Zhao,
`Stefan Gross,
`Lenny Dang,
`Katharine E. Yen,
`Hua Yang,
`Kimberly S. Straley,
`†
`†
`∇
`#
`#
`#
`Shengfang Jin,
`Kaiko Kunii,
`Valeria R. Fantin,
`Shunan Zhang,
`Qiongqun Pan,
`Derek Shi,
`†
`†
`Scott A. Biller,
`and Shinsan M. Su
`†
`Agios Pharmaceuticals, 38 Sidney Street, Cambridge, Massachusetts 02139, United States
`‡
`Ember Therapeutics, 855 Boylston Street, 11th Floor, Suite B, Boston, Massachusetts 02116, United States
`§Sage Therapeutics, 215 First Street, Cambridge, Massachusetts 02141, United States
`∥
`Schrödinger, Inc., 120 West 45th Street, New York, New York 10036, United States
`⊥
`Sundia MediTech Company, Ltd., Building 8, 388 Jialilue Road, Zhangjiang High-Tech Park, Shanghai 201203, China
`∇
`Oncology Research Unit, Pfizer Worldwide Research and Development, La Jolla Laboratories, San Diego, California 92121, United
`States
`#Shanghai ChemPartner Co., LTD, 998 Halei Road, Zhangjiang Hi-tech Park, Pudong New Area, Shanghai 201203, China
`*S Supporting Information
`
`ABSTRACT: Optimization of a series of R132H IDH1
`inhibitors from a high throughput screen led to the first
`potent molecules that show robust tumor 2-HG inhibition in a
`xenograft model. Compound 35 shows good potency in the
`U87 R132H cell based assay and ∼90% tumor 2-HG inhibition
`in the corresponding mouse xenograft model following BID
`dosing. The magnitude and duration of tumor 2-HG inhibition
`correlates with free plasma concentration.
`
`KEYWORDS: Mutant IDH1, tumor 2-HG, R132H IDH1 inhibitors
`
`T he family of isocitrate dehydrogenases (IDHs) includes
`
`two NADP dependent isoforms IDH1 and IDH2, which
`catalyze the oxidative decarboxylation of isocitrate to produce
`carbon dioxide, α-ketoglutarate (α-KG), and NADPH.1,2,14
`The implication of a role for IDH in cancer was revealed after
`somatic mutations in IDH1 were identified through a genome
`wide mutation analysis in glioblastoma.3 This landmark study
`was followed by high throughput sequencing, which revealed
`the presence of mutations in IDH1 in more than 70% of grade
`II−III gliomas and secondary glioblastomas,4 as well as in
`approximately 10−15% of patients with acute myeloid leukemia
`(AML).5 These somatic mutations were found at a key arginine
`residue belonging to the catalytic triad found in the enzyme’s
`active site (R132 for IDH1). This active site mutation results in
`loss-of-function for the oxidative decarboxylation of isocitrate
`and confers a novel gain-of-function for the production of the
`oncometabolite D-2-hydroxyglutarate (2-HG).6 Further charac-
`terization of
`the mutation showed that overexpression of
`mutant IDH1 in U87-MG, a human glioblastoma cell
`line,
`resulted in 100-fold elevated levels of 2-HG relative to the same
`cells expressing vector alone (data not shown).6 Recently, it
`was demonstrated that 2-HG is a competitive inhibitor of
`multiple α-KG-dependent dioxygenases, including histone and
`
`DNA demethylases,7,8 and several studies have shown that 2-
`HG producing IDH mutants are involved in global histone and
`DNA methylation alterations which may contribute to
`tumorigenesis through epigenetic rewiring.9,10 Taken together,
`these findings implicate mutant IDH1 as an oncogene and a
`compelling drug target for new therapies for glioma and AML
`patients.
`In order to identify small molecule inhibitors of IDH1,11,12
`we conducted a high-throughput screening (HTS) campaign
`against R132H IDH1 mutant protein homodimer. Library
`screen followed by confirmation of the active hits provided
`phenyl-glycine inhibitor 1. Detailed kinetic mechanism-of-
`action studies showed compound 1 binding to be reversible and
`behaving as competitive inhibitor with respect to α-KG and
`uncompetitive with respect to NADPH (data not shown).
`Given its attractive chemical
`structure and well-defined
`inhibitory properties, we selected this compound as a starting
`point for further optimization.
`
`Received: August 3, 2012
`Accepted: September 1, 2012
`Published: September 17, 2012
`
`© 2012 American Chemical Society
`
`850
`
`dx.doi.org/10.1021/ml300225h | ACS Med. Chem. Lett. 2012, 3, 850−855
`
`Rigel Exhibit 1011
`Page 1 of 6
`
`

`

`ACS Medicinal Chemistry Letters
`
`Letter
`
`Figure 1. HTS hit 1 and phenyl-glycine scaffold synthesis.
`
`Figure 2. Key structural elements that influence the binding affinity of the phenyl-glycine scaffold.
`
`Table 1. C-Terminus R1 SAR
`
`aThe IC50 values for the R132H homodimer are the mean of at least two determinations performed as described in the Supporting Information.
`
`We report herein that optimization of 1 led to the
`identification of 35, the first reported R132H IDH1 inhibitor
`to show robust in vivo reduction of 2-HG levels in a tumor
`xenograft model.
`The phenyl-glycine scaffold was readily assembled via four
`component Ugi reaction,13 as depicted retrosynthetically in
`Figure 1. Compound 1 was synthesized using cyclopentyl
`isocyanide, o-methyl benzaldehyde, m-fluoroaniline, and (2-
`thiophen-2-yl) acetic acid as starting materials. If 2-chloroacetic
`acid is used for the Ugi acid component, intermediate 2 (R4 =
`Cl) is
`readily obtained and can be used for
`further
`functionalization at R4 through nucleophilic displacement of
`the chlorine.
`Upon identification of 1 as a screening hit, we set out to
`understand the key structural elements that were responsible
`for the binding affinity of this compound to the R132H IDH1
`
`protein (Figure 2). The molecule displays mostly hydrophobic
`features, with three aromatic rings positioned around two
`amide carbonyl groups, with a rather high clogP (5.6). Starting
`from a closely related analog 3 (IC50 = 0.08 μM), we first
`initiated a substitution pattern investigation of the phenyl-
`glycine backbone. The eutomer/distomer relationship of the α-
`carbon stereocenter was established by chiral synthesis of
`analog 3 starting from D-and L-mandelic acid,14 which provided
`4 (S) and 5 (R) enantiomers, respectively, with compound 4
`(IC50 = 0.06 μM) possessing essentially all of the activity found
`in the racemate. The enantiospecificity of
`this enzyme
`inhibition held true in many analogs subsequently investigated
`(data not shown).
`rapid exploration of structure−activity relationships
`For
`(SARs), all subsequent compounds were profiled in their
`racemic form. Geminal substitution at the α-carbon as depicted
`
`851
`
`dx.doi.org/10.1021/ml300225h | ACS Med. Chem. Lett. 2012, 3, 850−855
`
`Rigel Exhibit 1011
`Page 2 of 6
`
`

`

`ACS Medicinal Chemistry Letters
`
`Table 2. N-Terminus R4 SAR
`
`Letter
`
`aThe IC50 values for R132H homodimer are the mean of at least two determinations performed as described in the Supporting Information.
`
`for 6 incurred an 18-fold potency loss compared to the case of
`3. Next, alkylation of the secondary amide nitrogen as shown
`for 7 caused a 45-fold loss in potency compared to the case of
`3, while replacement of either the C-terminus or N-terminus
`carbonyl groups (compounds 8 and 9) with a CH2 moiety
`resulted in significant loss of biochemical activity, highlighting
`the importance of both amide moieties for binding affinity.
`
`We then started a systematic investigation of SAR for the N-
`and C-terminus regions of the scaffold, as well as the central
`aromatic moieties, with a key objective to improve properties,
`including decreasing the lipophilicity of the initial hit 1, while
`improving biochemical potency.
`R1 functional group exploration (Table 1) revealed that
`carbocycles were well tolerated, with cyclohexyl 10 slightly
`better (IC50 = 0.05 μM) than the starting HTS hit 1. As the
`dx.doi.org/10.1021/ml300225h | ACS Med. Chem. Lett. 2012, 3, 850−855
`
`852
`
`Rigel Exhibit 1011
`Page 3 of 6
`
`

`

`ACS Medicinal Chemistry Letters
`
`Letter
`
`Table 3. Selectivity and Cell Based Profiling of Potent Phenyl-Glycine Analogs
`
`HT1080
`U87
`IDH1wt
`HT1080
`R132C
`U87
`R132H
`IC50 (μM)b
`GI50 (μM)
`IC50 (μM)
`IC50 (μM)a
`GI50 (μM)
`IC50 (μM)
`IC50(μM)a
`cLogP
`compd
`7.7 (32%)
`>20
`0.39
`0.05
`>20
`0.58
`0.09
`5.6
`1
`2.9 (22%)
`>20
`0.15
`0.03
`>20
`0.19
`0.05
`6.2
`10
`19.7 (32%)
`>20
`0.22
`0.06
`>20
`0.29
`0.10
`4.9
`18
`19.3 (34%)
`>20
`0.26
`0.05
`>20
`0.21
`0.05
`5.0
`19
`6.3 (39%)
`>3
`0.09
`0.03
`>20
`0.36
`0.06
`6.5
`20
`>100
`>20
`0.48
`0.16
`>20
`0.07
`0.07
`4.7
`35
`>100
`>20
`0.11
`0.04
`>20
`0.24
`0.08
`6.3
`36
`>100
`>20
`0.17
`0.04
`>20
`0.37
`0.07
`7.1
`37
`aThe IC50 values for R132H and R132C and wt homodimers are the mean of at least two determinations performed as described in the Supporting
`Information. bFor compounds with less than 100% enzyme inhibition, the maximum inhibition achieved is shown.
`
`Figure 3. Tumor 2-HG inhibition following one and three BID doses of 150 mg/kg of 35 via IP route in the U87 R132H tumor xenograft model.
`
`ring size decreased from cyclohexyl 10 to cyclopropyl 13,
`potency decreased gradually to low micromolar values.
`Replacement of cyclohexyl with aromatic rings (o-tolyl, benzyl)
`as shown for 14 and 15 led to a 10−30-fold decrease in
`biochemical potency compared to the case of 10. In an attempt
`to improve the properties of these compounds by decreasing
`the clogP through heteroatom substitution in the cyclohexyl
`ring, we found that pyran 16 was 10-fold less potent than 10,
`while piperidine 17 suffered a nearly 100-fold loss of
`biochemical potency.
`Evaluation of R2 substituents revealed that for the α-aromatic
`ring ortho-substitution was most favored, while replacement of
`the phenyl group with heterocycles or carbocycles afforded only
`low micromolar potency analogs.14 A preliminary survey of R3
`pointed to meta-substituted aromatic groups being most
`favorable, while aliphatic moieties, acyclic or cyclic, or
`heteroatom containing carbocycles provided analogs with a
`significant drop in biochemical potency.14 These observations
`coupled with the C-terminus amide SAR results shown in Table
`1 suggested that the phenyl-glycine scaffold was binding in a
`highly lipophilic region of the enzyme.
`Having elucidated SAR on three areas of the scaffold, we
`continued our exploration on the N-terminus R4 substituents in
`an additional approach to improve the compound physical
`chemical properties and decrease the overall lipophilicity of the
`original hit. Synthetic chemistry readily amenable to parallel
`arrays allowed us to rapidly explore a variety of functional
`groups at the N-terminus (Table 2).
`
`Replacement of thiophene in compound 1 by other carbon-
`linked heterocycles, such as thiazole 18, 4-pyridyl 19, or 3-
`indole 20 analogs, provided similar biochemical potency to 1 in
`the 0.05−0.1 μM range. Modification of the pyridine 19 to
`pyrimidine 21 caused a 22-fold drop in potency. Our
`investigation next focused on nitrogen linked systems directly
`prepared from intermediate 2 (R4 = Cl) via chlorine
`displacement (Figure 1). Replacement of chlorine in compound
`22 with amino group in analog 23 caused a potency drop from
`0.9 to 7.8 μM. A small set of cycloalkyl amines with an
`adjustment of ring size led to diminishing biochemical potency
`from cyclopropyl 24 (0.19 μM) to cyclohexyl 26 (1.27 μM).
`Interestingly, when the cyclohexyl group in 26 was replaced by
`a phenyl ring in compound 27, biochemical potency was
`substantially enhanced from 1.27 μM to 0.06 μM. Alkylation of
`the aniline nitrogen as shown in analog 28 maintained the
`potency at 0.05 μM. Use of aromatic rings for R4 did improve
`the potency compared to the aliphatic analogs (24−26);
`however, clogP increased as well,
`retaining their highly
`hydrophobic character. Encouraged by the good potency of
`aromatic tertiary amine 28, we next tested a small set of tertiary
`aliphatic amines, with the aim of decreasing the lipophilicity of
`the scaffold by introduction of basic solubilizing groups. Among
`the examples that were evaluated, morpholine 30 displayed the
`best biochemical potency (0.24 μM), while pyrrolidine 29 and
`piperazine 31 afforded weakly active single digit micromolar
`analogs. Addition of a fused phenyl ring to pyrrolidine as shown
`in 32 improved the potency by 10-fold at the expense of an
`
`853
`
`dx.doi.org/10.1021/ml300225h | ACS Med. Chem. Lett. 2012, 3, 850−855
`
`Rigel Exhibit 1011
`Page 4 of 6
`
`

`

`ACS Medicinal Chemistry Letters
`
`increased clogP value. We then continued exploration of
`nitrogen heterocycles from intermediate 2. Replacing the
`pyrrolidine ring in compound 29 with pyrazole 33 improved
`biochemical potency by 4-fold, while triazole 34 offered a slight
`improvement in binding affinity. Gratifyingly, N-(2-methyl)-
`imidazole 35 restored the biochemical potency to 0.07 μM,
`while fused nitrogen linked heterocycles (benzimidazole 26,
`indole 37) maintained the same potency at the expense of
`increased clogP values. Overall, R4-substitution allowed
`introduction of a diverse set of substituents that were well
`tolerated, possibly indicating that this part of the phenyl glycine
`scaffold may be binding in a solvent exposed area of the
`protein.
`As
`the SAR investigation revealed functional group
`modifications that provided potent inhibitors in the R132H
`enzymatic assay, we selected a focused set of analogs for
`evaluation against R132C IDH1 mutant15 and wild-type IDH1
`enzymes. Additionally, compounds were profiled in the
`glioblastoma U87 cells
`that overexpress mutant R132H
`IDH1, as well as the HT1080 chondrosarcoma cell
`line,
`which expresses the endogenous R132C IDH1 mutant.16 These
`cell lines produce significant levels of 2-HG compared to vector
`cells alone. Upon treatment with inhibitor for 48 h, the levels of
`2-HG were measured in the media by LCMS, to generate IC50
`values.14 Within the same experiment, 50% growth inhibition
`(GI50) was determined by measuring total cellular ATP after 72
`h of compound treatment.
`As shown in Table 3, the majority of compounds showed
`similar biochemical potency against the R132C IDH1 mutant
`and displayed cellular IC50 values less than 0.5 μM in both U87
`and HT1080 cell lines, with a 3−5-fold shift in enzyme to cell
`potency in most cases. Exquisite selectivity for R132H and
`R132C IDH1 mutant isoforms was demonstrated by the poor
`biochemical activity against the wild-type IDH1 and the lack of
`induction of nonspecific cell death (GI50 > 20 μM).
`Compound 35, equipotent in both enzyme R132H and U87
`cellular assays, was selected for additional in vivo profiling in the
`U87 R132H tumor xenograft mouse model (Figure 3). In vitro
`and in vivo DMPK studies were conducted for compound 35.
`This analog showed rapid turnover
`in human and rat
`microsomal incubations with an estimated hepatic extraction
`ratio of 0.93 and 0.85, respectively. Plasma protein binding was
`95.7% in mouse using the equilibrium dialysis method.
`Reasonable plasma exposure was achieved via intraperitoneal
`dosing at 50 mg/kg (AUC0−24h = 20800 h·ng/mL), enabling
`the use of inhibitor 35 for further in vivo studies. Female nude
`mice bearing U87 R132H tumor xenografts14 were dosed via IP
`route with 150 mg/kg of 35 formulated in 0.5% MC and 0.2%
`Tween 80, and then they were compared to the vehicle control
`animals. Blood and tumor samples were taken at different time
`points following compound administration. The plasma and
`inhibitor 35, as well as
`tumor concentrations of
`the
`corresponding tumor 2-HG concentrations were determined
`using sensitive and specific LC/MS/MS methods. The
`unbound plasma concentration of 35 was calculated using the
`total plasma concentration of 35 and free fraction of 35 in
`mouse plasma (4.3%).
`Following a single dose of 35, the estimated plasma free
`concentration of 35 was higher than the in vitro cellular IC50
`value (0.07 μM) for over 10 h. The magnitude and duration of
`tumor 2-HG inhibition correlated well with the free plasma
`concentration of 35. Compared to a single dose, a repeat dose
`of 35 provided longer exposure coverage time (drug exposure >
`
`Letter
`
`IC50) while the Cmax of 35 was similar following single and BID
`dosing. Better tumor 2-HG inhibition was achieved following
`BID dosing compared to a single dose, where the maximum
`tumor 2-HG inhibition was 89.4% and 69%, respectively. These
`results demonstrated that tumor 2-HG inhibition correlated
`with the duration of drug exposure and that robust tumor 2-HG
`inhibition is achievable with adequate and sustainable drug
`exposure.
`In conclusion, we have discovered the first class of potent
`IDH1 mutant inhibitors through optimization of HTS hits.
`Compound 35 is a potent inhibitor of 2-HG production in U87
`R132H cells and shows ∼90% tumor 2-HG inhibition in vivo
`following three BID doses. As high levels of 2-HG have been
`shown to alter the epigenetic state and biology of cells,9,10,17 the
`utility of this molecule will be important to assess the biological
`consequences of IDH mutations and the potential of IDH
`inhibitors for treating IDH mutant tumors.
`
`■ ASSOCIATED CONTENT
`*S Supporting Information
`in vivo studies,
`Experimental procedures for assay protocols,
`and synthesis and characterization of compounds. This material
`is available free of charge via the Internet at http://pubs.acs.org.
`
`■ AUTHOR INFORMATION
`Corresponding Author
`*Tel: (617) 649-8604. Fax: (617) 649-8618. E-mail: janeta.
`popovici-muller@agios.com.
`Notes
`The authors declare no competing financial interest.
`
`■ ACKNOWLEDGMENTS
`
`We thank Dr. Nageshwara Rao KV and Dr. Sarma BVNBS at
`SAI Advantium for their contribution to the synthesis of
`compound 8.
`
`■ REFERENCES
`
`(1) Yen, K. E.; Bittinger, M. A.; Su, S. M.; Fantin, V. R.; Cancer-
`associated, I. D. H. Mutations: biomarker and therapeutic oppor-
`tunities. Oncogene 2010, 29, 6409−6417.
`(2) Reitman, Z. J.; Yan, H. Isocitrate dehydrogenase 1 and 2
`mutations in cancer: alterations at a crossroads of cellular metabolism.
`J. Natl. Cancer Inst. 2010, 102, 932−941.
`(3) Parsons, D. W.; Jones, S.; Zhang, X.; Lin, J. C.; Leary, R. J.;
`Angenendt, P.; Mankoo, P.; Carter, H.; Siu, I. M.; Gallia, G. L.; Olivi,
`A.; McLendon, R.; Rasheed, B. A.; Keir, S.; Nikolskaya, T.; Nikolsky,
`Y.; Busam, D. A.; Tekleab, H.; Diaz, L. A., Jr.; Hartigan, J.; Smith, D.
`R.; Strausberg, R. L.; Marie, S. K.; Shinjo, S. M.; Yan, H.; Riggins, G. J.;
`Bigner, D. D.; Karchin, R.; Papadopoulos, N.; Parmigiani, G.;
`Vogelstein, B.; Velculescu, V. E.; Kinzler, K. W. An integrated
`genomic analysis of human glioblastoma multiforme. Science 2008,
`321, 1807−1812.
`(4) Yan, H.; Parsons, D. W.; Jin, G.; McLendon, R.; Rasheed, B. A.;
`Yuan, W.; Kos, I.; Batinic-Haberle, I.; Jones, S.; Riggins, G. J.;
`Friedman, H.; Friedman, A.; Reardon, D.; Herndon, J.; Kinzler, K. W.;
`Velculescu, V. E.; Vogelstein, B.; Bigner, D. D. IDH1 and IDH2
`mutations in gliomas. N. Engl. J. Med. 2009, 360, 765−73.
`(5) Paschka, P.; Schlenk, R. F.; Gaidzik, V. I.; Habdank, M.; Krönke,
`J.; Bollinger, L.; Späth, D.; Kayser, S.; Zucknick, M.; Götze, K.; Horst,
`H. A.; Germing, U.; Döhner, H.; Döhner, K. IDH1 and IDH2
`mutations are frequent genetic alterations in acute myeloid leukemia
`and confer adverse prognosis in cytogenetically normal acute myeloid
`leukemia with NPM1 mutation without FLT3 internal
`tandem
`duplication. J. Clin. Oncol. 2010, 28, 3636−3643.
`
`854
`
`dx.doi.org/10.1021/ml300225h | ACS Med. Chem. Lett. 2012, 3, 850−855
`
`Rigel Exhibit 1011
`Page 5 of 6
`
`

`

`Letter
`
`ACS Medicinal Chemistry Letters
`
`(6) Dang, L.; White, D. W.; Gross, S.; Bennett, B. D.; Bittinger, M.
`A.; Driggers, E. M.; Fantin, V. R.; Jang, H. G.; Jin, S.; Keenan, M. C.;
`Marks, K. M.; Prins, R. M.; Ward, P. S.; Yen, K. E.; Liau, L. M.;
`Rabinowitz, J. D.; Cantley, L. C.; Thompson, C. B.; Vander Heiden,
`M. G.; Su, S. M. Cancer-associated IDH1 mutations produce 2-
`hydroxyglutarate. Nature 2009, 462, 739−44.
`(7) Xu, W.; Yang, H.; Liu, Y.; Yang, Y.; Wang, P.; Kim, S. H.; Ito, S.;
`Yang, C.; Wang, P.; Xiao, M. T.; Liu, L. X.; Jiang, W. Q.; Liu, J.; Zhang,
`J. Y.; Wang, B.; Frye, S.; Zhang, Y.; Xu, Y. H.; Lei, Q. Y.; Guan, K. L.;
`Zhao, S. M.; Xiong, Y. Oncometabolite 2-hydroxyglutarate is a
`competitive inhibitor of α-ketoglutarate-dependent dioxygenases.
`Cancer Cell 2011, 19, 17−30.
`(8) Chowdhury, R.; Yeoh, K. K.; Tian, Y. M.; Hillringhaus, L.; Bagg,
`E. A.; Rose, N. R.; Leung, I. K.; Li, X. S.; Woon, E. C.; Yang, M.;
`McDonough, M. A.; King, O. N.; Clifton, I. J.; Klose, R. J.; Claridge, T.
`D.; Ratcliffe, P. J.; Schofield, C. J.; Kawamura, A. The oncometabolite
`2-hydroxyglutarate inhibits histone lysine demethylases. EMBO Rep.
`2011, 12, 463−469.
`(9) Lu, C.; Ward, P. S.; Kapoor, G. S.; Rohle, D.; Turcan, S.; Abdel-
`Wahab, O.; Edwards, C. R.; Khanin, R.; Figueroa, M. E.; Melnick, A.;
`Wellen, K. E.; O’Rourke, D. M.; Berger, S. L.; Chan, T. A.; Levine, R.
`L.; Mellinghoff, I. K.; Thompson, C. B. IDH mutation impairs histone
`demethylation and results in a block to cell differentiation. Nature
`2012, 483, 474−478.
`(10) Turcan, S.; Rohle, D.; Goenka, A.; Walsh, L. A.; Fang, F.;
`Yilmaz, E.; Campos, C.; Fabius, A. W.; Lu, C.; Ward, P. S.; Thompson,
`C. B.; Kaufman, A.; Guryanova, O.; Levine, R.; Heguy, A.; Viale, A.;
`Morris, L. G.; Huse, J. T.; Mellinghoff, I. K.; Chan, T. A. IDH1
`mutation is sufficient
`to establish the glioma hypermethylator
`phenotype. Nature 2012, 483, 474−478.
`(11) Popovici-Muller, J.; Salituro, F. G.; Saunders, J. O.; Travins, J.;
`Yan, S. Preparation of cycloalkylheteroarylacetylphenylaminophenyla-
`cetamide derivatives and analogs for use as antitumor agents. WO
`2012009678.
`(12) Salituro, F. G.; Saunders, J. O. Preparation of piperazinylcar-
`bonyl benzenesulfonamides for use in the treatment of cancer
`characterized as having an IDH mutation. WO 2011072174.
`(13) Ugi, I.; Meyr, R.; Fetzer, U.; Steinbrückner, C. Versuche mit
`Isonitrilen. Angew. Chem. 1959, 71 (11), 386. Ugi, I.; Lohberger, S.;
`Karl, R. The Passerini and Ugi Reactions, Chapter 4.6. Comprehensive
`Organic Synthesis 1991, 2, 1083−1109.
`(14) See Supporting Information.
`(15) Hartmann, C.; Meyer, J.; Balss, J.; Cappe, D.; Mueller, W.;
`Christians, A.; Felsberg, J.; Wolter, M.; Mawrin, C.; Wick, W.; Weller,
`M.; Herold-Mende, C.; Unterberg, A.; Jeuken, J. W.; Wesseling, P.;
`Reifenberger, G.; von Deimling, A. Type and frequency of IDH1 and
`IDH2 mutations are related to astrocytic and oligodendroglial
`differentiation and age: a study of 1,010 diffuse gliomas. Acta
`Neuropathol. 2009, 118, 469−474.
`(16) Amary, M. F.; Bacsi, K.; Maggiani, F.; Damato, S.; Halai, D.;
`Berisha, F.; Pollock, R.; O’Donnell, P.; Grigoriadis, A.; Diss, T.;
`Eskandarpour, M.; Presneau, N.; Hogendoorn, P. C.; Futreal, A.;
`Tirabosco, R.; Flanagan, A. M. IDH1 and IDH2 mutations are
`frequent events in central chondrosarcoma and central and periosteal
`chondromas but not in other mesenchymal tumours. J. Pathol. 2011,
`224, 334−343.
`(17) Sasaki, M.; Knobbe, C. B.; Munger, J. C.; Lind, E, F.; Brenner,
`D.; Brüstle, A.; Harris, I. S.; Holmes, R.; Wakeham, A.; Haight, J.; You-
`Ten, A.; Li, W. Y.; Schalm, S.; Su, S. M.; Virtanen, C.; Reifenberger,
`G.; Ohashi, P. S.; Barber, D. L.; Figueroa, M. E.; Melnick, A.; Zúñiga-
`Pflücker, J. C.; Mak, T. W. IDH1(R132H) mutation increases murine
`haematopoietic progenitors and alters epigenetics. Nature 2012, 488,
`656−659.
`
`
`
`View publication statsView publication stats
`
`855
`
`dx.doi.org/10.1021/ml300225h | ACS Med. Chem. Lett. 2012, 3, 850−855
`
`Rigel Exhibit 1011
`Page 6 of 6
`
`

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