throbber
REPORTS
`
`30 hours after infection of mosquitoes lacking
`LRIM1, APL1, or TEP1. As reported previously
`(16), three distinct classes of parasites were ob-
`served in the midguts of control mosquitoes: live
`(GFP positive), dead (TEP1 positive), and dying
`(GFP and TEP1 positive). In contrast, TEP1-positive
`parasites were never detected in LRIM1 or APL1
`KD mosquitoes (Fig. 3C), despite the presence
`of TEP1-F in the hemolymph. Lack of TEP1 par-
`asite staining was as complete as in TEP1 KD
`mosquitoes, which entirely lack TEP1 in the hemo-
`lymph. These data demonstrate that the LRIM1/
`APL1C complex is necessary for TEP1-mediated
`parasite killing during midgut invasion and in-
`dicate that TEP1 binds parasites only after it is
`processed.
`The APL1 locus has been implicated in
`mosquito resistance to the human malaria para-
`site, Plasmodium falciparum (6), and TEP1 has
`been shown to act against P. falciparum in labo-
`ratory infections (7). Mosquito defense against
`Plasmodium is likely to be influenced by vector/
`parasite coevolution and adaptation; thus, the ob-
`servation that LRIM1 did not affect P. falciparum
`in experimental field infections (17) may suggest
`that parasites have evolved to evade this pathway.
`Proteins such as the fibrinogen-related FBN9 (18)
`or other LRR proteins may provide alternative
`mechanisms for TEP1-mediated parasite killing.
`Bioinformatic searches for proteins related to
`LRIM1 and APL1C using their shared structural
`features (signal peptide, LRRs, cysteine pattern,
`and coiled-coils) (see supporting online material)
`detected more than 20 LRIM-like genes in each
`of the available mosquito genomes—A. gambiae,
`Aedes aegypti, and Culex quinquefasciatus—but
`not in any other species (table S1). Several of these
`genes were previously implicated in A. gambiae
`immune responses (3, 6, 7, 9, 19). Phylogenetic
`analysis in conjunction with pairwise compar-
`isons, examination of orthologous genomic neigh-
`borhoods, and protein domain analysis revealed
`four distinct LRIM subfamilies (figs. S3 and S4).
`Thus, LRIM1 and APL1C are members of a
`family of putative recognition receptors, which
`appears to be unique and greatly expanded in
`mosquitoes. Nevertheless, structural integrity of
`both LRRs and coiled-coils rests with only a few
`key amino acids, allowing considerable sequence
`variation that may hinder identification of func-
`tional equivalents in other organisms.
`The versatile LRR motif mediates recognition
`of diverse pathogen-associated molecules in host
`innate defense in plants and animals (20). For
`example, the repertoire of variable lymphocyte
`receptor (VLR) antibodies in jawless vertebrates
`is generated by combinatorial assembly of
`LRR modules instead of immunoglobulin seg-
`ments as in jawed vertebrates (21). Similarly to
`LRIM1/APL1C, the VLR antibodies are secreted
`as disulfide-linked multimers (22).
`LRIMs form a family of mosquito LRR re-
`ceptors with putative roles in defense against hu-
`man and animal pathogens. LRIM1 and APL1C
`exist as a complex that mediates immunity
`
`against malaria parasites through activation of
`mosquito complement. The multimeric nature of
`the complex indicates the potential to bind mul-
`tiple targets similarly to mammalian multisubunit
`receptors that robustly activate complement, that
`is, immunoglobulin M, lectin, and C1q. Bound
`LRIM1/APL1C complex may then undergo con-
`formational changes inducing the recruitment of
`additional cascade components, such as a TEP1-
`activating protease. In-depth study of these in-
`teractions will provide insights into complement
`activation in mosquitoes and tools toward block-
`ing disease transmission.
`References and Notes
`1. R. W. Snow, C. A. Guerra, A. M. Noor, H. Y. Myint, S. I. Hay,
`Nature 434, 214 (2005).
`2. S. A. Blandin, E. Marois, E. A. Levashina, Cell Host
`Microbe 3, 364 (2008).
`3. M. A. Osta, G. K. Christophides, F. C. Kafatos, Science
`303, 2030 (2004).
`4. E. Warr, L. Lambrechts, J. C. Koella, C. Bourgouin,
`G. Dimopoulos, Insect Biochem. Mol. Biol. 36, 769
`(2006).
`5. L. F. Moita et al., Immunity 23, 65 (2005).
`6. M. M. Riehle et al., Science 312, 577 (2006).
`7. Y. Dong et al., PLoS Pathog. 2, e52 (2006).
`8. T. Habtewold, M. Povelones, A. M. Blagborough,
`G. K. Christophides, PLoS Pathog. 4, e1000070 (2008).
`9. M. M. Riehle et al., PLoS One 3, e3672 (2008).
`10. F. H. Collins et al., Science 234, 607 (1986).
`11. G. Dimopoulos, A. Richman, H. M. Muller, F. C. Kafatos,
`Proc. Natl. Acad. Sci. U.S.A. 94, 11508 (1997).
`12. H. M. Muller, G. Dimopoulos, C. Blass, F. C. Kafatos,
`J. Biol. Chem. 274, 11727 (1999).
`
`13. S. Blandin et al., Cell 116, 661 (2004).
`14. E. A. Levashina et al., Cell 104, 709 (2001).
`15. R. H. Baxter et al., Proc. Natl. Acad. Sci. U.S.A. 104,
`11615 (2007).
`16. C. Frolet, M. Thoma, S. Blandin, J. A. Hoffmann,
`E. A. Levashina, Immunity 25, 677 (2006).
`17. A. Cohuet et al., EMBO Rep. 7, 1285 (2006).
`18. Y. Dong, G. Dimopoulos, J. Biol. Chem. www.jbc.org/cgi/
`doi/10.1074/jbc.M807084200. Published online
`4 February 2009.
`19. R. Aguilar et al., Insect Biochem. Mol. Biol. 35, 709 (2005).
`20. T. Nurnberger, F. Brunner, B. Kemmerling, L. Piater,
`Immunol. Rev. 198, 249 (2004).
`21. Z. Pancer et al., Nature 430, 174 (2004).
`22. B. R. Herrin et al., Proc. Natl. Acad. Sci. U.S.A. 105,
`2040 (2008).
`23. The authors thank A. C. Koutsos for generating the LRIM1
`antibody and sharing it before publication and F. M. Ausubel
`for critically reviewing the manuscript. Fluorescence
`microscopy was performed at the Imperial College Facility for
`Imaging by Light Microscopy imaging facility. This work
`was supported by a Wellcome Trust Program grant
`(GR077229/Z/05/Z), an NIH Program Project
`(2PO1AI044220-06A1), and a Biotechnology and Biological
`Sciences Research Council grant (BB/E002641/1). R.M.W.
`was supported by a Wellcome Trust Ph.D. fellowship.
`Supporting Online Material
`www.sciencemag.org/cgi/content/full/1171400/DC1
`Materials and Methods
`Figs. S1 to S4
`Table S1
`References
`
`26 January 2009; accepted 24 February 2009
`Published online 5 March 2009;
`10.1126/science.1171400
`Include this information when citing this paper.
`
`Glioma-Derived Mutations in IDH1
`Dominantly Inhibit IDH1 Catalytic
`Activity and Induce HIF-1a
`Shimin Zhao,1,2 Yan Lin,1* Wei Xu,1,2* Wenqing Jiang,1,2* Zhengyu Zha,1 Pu Wang,1,2
`Wei Yu,1,2 Zhiqiang Li,4 Lingling Gong,5 Yingjie Peng,6 Jianping Ding,6 Qunying Lei,1,3
`Kun-Liang Guan,1,3,7† Yue Xiong1,2,8†
`Heterozygous mutations in the gene encoding isocitrate dehydrogenase-1 (IDH1) occur in certain
`human brain tumors, but their mechanistic role in tumor development is unknown. We have shown
`that tumor-derived IDH1 mutations impair the enzyme’s affinity for its substrate and dominantly
`inhibit wild-type IDH1 activity through the formation of catalytically inactive heterodimers.
`Forced expression of mutant IDH1 in cultured cells reduces formation of the enzyme product,
`a-ketoglutarate (a-KG), and increases the levels of hypoxia-inducible factor subunit HIF-1a, a
`transcription factor that facilitates tumor growth when oxygen is low and whose stability is
`regulated by a-KG. The rise in HIF-1a levels was reversible by an a-KG derivative. HIF-1a levels
`were higher in human gliomas harboring an IDH1 mutation than in tumors without a mutation.
`Thus, IDH1 appears to function as a tumor suppressor that, when mutationally inactivated,
`contributes to tumorigenesis in part through induction of the HIF-1 pathway.
`
`Gliomas are the most common type of
`
`human brain tumors and can be classified
`based on clinical and pathological criteria
`into four grades. The grade IV glioma, commonly
`known as glioblastoma multiforme (GBM), has
`one of the worst prognoses among all types of
`human tumors and can develop either de novo
`(primary GBM) or through progression from low-
`grade tumors (secondary GBM). Although patho-
`
`logically indistinguishable, primary and secondary
`GBM exhibit distinct patterns of cancer gene al-
`terations (1). A recent cancer genome sequencing
`project revealed that the gene encoding IDH1 is
`somatically mutated predominantly in secondary
`GBM (2). Three subsequent studies of targeted
`IDH1 gene sequencing confirmed this finding,
`together identifying IDH1 mutations in more than
`70% of secondary GBM or low-grade gliomas but
`
`www.sciencemag.org SCIENCE VOL 324
`
`10 APRIL 2009
`
`261
`
`Rigel Exhibit 1012
`Page 1 of 5
`
`

`

`REPORTS
`
`infrequently in primary GBM (about 5%) (3–5).
`Notably, all of the IDH1 mutations identified to
`date produce a single amino acid substitution at
`Arg132 (R132) and no obvious inactivating (frame-
`shift or protein-truncation) mutations were found.
`This observation, together with the fact that the
`tumors do not show loss-of-heterozygosity (LOH),
`has led to speculation that R132 mutations lead to
`oncogenic activation of the enzyme.
`IDH enzymes catalyze the oxidative decar-
`boxylation of isocitrate (ICT) to produce a-KG.
`The human genome has five IDH genes coding
`for three distinct IDH enzymes whose activities
`are dependent on either nicotinamide adenine di-
`nucleotide phosphate (NADP+-dependent IDH1
`and IDH2) or nicotinamide adenine dinucleo-
`tide (NAD+-dependent IDH3). Both IDH2 and
`IDH3 enzymes are localized in the mitochon-
`dria and participate in the citric acid (TCA) cycle
`for energy production, whereas IDH1 is localized
`in the cytoplasm and peroxisomes (6). The R132
`residue is conserved in all NADP+-dependent
`IDHs.
`To explore the functional impact of the tumor-
`associated mutations at R132, we performed mod-
`eling studies based on the previously reported
`human cytosolic IDH1 crystal structure (7).
`Among all residues involved in binding with
`ICT, the side chain of R132 uniquely forms three
`hydrogen bonds with both the a- and b-carboxyl
`groups of the substrate ICT, whereas other resi-
`dues involved in ICT binding form no more than
`two hydrogen bonds (Fig. 1A). Substitution of
`R132 with any one of the six amino acids ob-
`served in gliomas (His, Ser, Gly, Cys, Val, and
`Leu) would impair interactions of the enzyme with
`ICT both sterically and electrostatically. Repre-
`sentative modeling of H132, which corresponds to
`the most prevalent IDH1 mutation in human
`gliomas, is shown in Fig. 1A. We determined the
`in vitro enzymatic activities of three tumor-derived
`IDH1 mutants, R132H, R132C and R132S, ex-
`pressed in transformed human embryonic kidney
`(HEK) 293T cells and found that all three mutants
`have a greater than 80% reduction in activity
`as compared with the wild-type (WT) IDH1
`(Fig. 1B). Analysis of recombinant IDH1 mutant
`
`1Molecular and Cell Biology Laboratory, Institute of Biomedical
`Sciences, Fudan University, 130 Dong-An Road, Shanghai
`200032, China. 2School of Life Sciences, Fudan University, 220
`Han-Dan Road, Shanghai 200433, China. 3Department of
`Biological Chemistry, School of Medicine, Fudan University,
`130 Dong-An Road, Shanghai 200032, China. 4Department of
`Neurosurgery, Zhongnan Hospital, Wuhan University, Wuhan
`430071, China. 5Department of Pathology, Zhongnan
`Hospital, Wuhan University, Wuhan 430071, China. 6State
`Key Laboratory of Molecular Biology, Institute of Biochemistry
`and Cell Biology, Shanghai Institute for Biological Sciences,
`Chinese Academy of Sciences, 320 Yue-Yang Road, Shanghai
`200031, China. 7Department of Pharmacology and Moores
`Cancer Center, University of California San Diego, La Jolla, CA
`92093, USA. 8Department of Biochemistry and Biophysics,
`Lineberger Comprehensive Cancer Center, University of North
`Carolina at Chapel Hill, NC 27599, USA.
`*These authors contributed equally to this work.
`†To whom correspondence should be addressed. E-mail:
`kuguan@ucsd.edu (K.-L.G.); yxiong@email.unc.edu (Y.X.)
`
`proteins purified from Escherichia coli likewise
`displayed little activity in vitro compared with
`wild-type controls (Fig. 1B). Kinetic analyses of
`recombinant IDH1 proteins revealed that all three
`mutant IDH1s had a dramatically reduced affini-
`ty for ICT: The Michaelis constants (Kms) of
`IDH1R132C, IDH1R132S, and IDH1R132H for ICT
`were increased by factors of 60, 70, and 94, re-
`spectively (Fig. 1C). In contrast, the Km for
`NADP+ and the maximum velocity (Vmax) were
`not appreciably altered (Fig. 1C). Similarly,
`mutation of an arginine residue in pig mitochon-
`drial IDH2 equivalent to R132 in human IDH1
`caused a dramatic increase in Km for isocitrate
`(by a factor of 165), with minimal effect on Vmax
`(8). Because the normal cellular concentration
`of ICT is 20 to 30 mM (9), which is lower than
`the Km of the IDH1 mutants, the mutant en-
`zymes are likely to have limited activity under
`physiological conditions. Together, these struc-
`tural and biochemical analyses indicate that
`the tumor-associated IDH1 mutations inactivate
`the enzyme.
`Given that IDH1 normally functions as a
`homodimer, we hypothesized that the mutant
`IDH1 molecules in tumor cells form heterodimers
`with wild-type molecules and, in so doing, domi-
`nantly inhibit the activity of wild-type IDH1. To
`test this hypothesis, we coexpressed His-tagged
`wild-type and FLAG-tagged R132H mutant
`IDH1 in E. coli and isolated the heterodimer by
`
`sequential affinity purification using first nickel
`resin and then FLAG beads. Formation of either
`WT:WT homodimer or WT:R132H heterodimer
`was confirmed by gel filtration (fig. S1). As ex-
`pected,
`the WT:WT homodimers were fully
`active and the R132H:R132H homodimers were
`nearly completely inactive (Fig. 2A). Notably, the
`WT:R132H heterodimer exhibited only 4% of the
`activity shown by the wild-type enzyme when
`assayed with limited ICT concentration (Fig. 2A).
`Normally, IDH1 can adopt at least three distinct
`conformations during catalysis: a quasi-open con-
`formation when it is in a complex with NADP+, a
`quasi-closed conformation when it is in a com-
`plex with ICT, and a closed conformation when it
`is in a complex with both NADP+ and ICT (fig.
`S2A) (7). The two IDH1 subunits act in a co-
`operative manner and undergo conformational
`changes in a concerted way. Our modeling study
`suggests that the impairment in enzyme binding
`with ICT conferred by the R132 mutation in one
`subunit might also impair the binding of ICT to
`the second wild-type subunit. As a result, both
`subunits would be locked in an unliganded or
`quasi-open (NADP+-bound) conformation,
`thereby inhibiting catalytic activity (Fig. 2B for
`the close-up of the catalytic active site) (fig. S2,
`B and C). Consistent with this model, we found
`that the wild-type IDH1 exhibits a sigmoidal
`curve of cooperative binding to ICT, whereas
`the WT:R132H heterodimer displayed a hy-
`
`Fig. 1. Tumor-derived IDH1 mutants have reduced catalytic activity because of impaired isocitrate
`binding. (A) Structural modeling predicts that mutation of R132 in IDH1 would weaken hydrogen bonding
`of the enzyme to ICT. Shown is a view of the catalytic active site of human IDH1 bound with NADP+
`(omitted for clarity), ICT (green), and Ca2+. The residues interacting with ICT from the adjacent subunit are
`labeled with an apostrophe. Hydrogen-bonding interactions are indicated with dashed lines. Simulated
`H132 mutation (cyan) is superimposed on R132. (B) Tumor-derived IDH1 mutants have reduced catalytic
`activity in vitro. Left, FLAG-tagged wild-type and mutant IDH1 were expressed in HEK293T cells, purified
`by immunoprecipitation and eluted by FLAG peptide; right, HIS-tagged wild-type and mutant IDH1 were
`expressed in E. coli and purified by nickel resin. Specific IDH1 activities for all proteins were measured in
`the presence of NADP+ (10 mM) and ICT (30 mM), with the presence of 2 mM Mn2+. Shown are mean
`values of triplicate experiments TSD. (C) Kinetic parameters of wild-type and mutant IDH1. Shown are
`mean values of duplicate experiments TSD.
`
`262
`
`10 APRIL 2009 VOL 324 SCIENCE www.sciencemag.org
`
`Rigel Exhibit 1012
`Page 2 of 5
`
`

`

`perbolic curve with a much higher Km (Fig. 2C),
`indicating that the heterodimer not only loses
`affinity but also cooperativity toward ICT.
`We next investigated whether loss of IDH1
`activity would alter cellular levels of a-KG, the
`product of IDH1 catalysis. We used RNA in-
`terference to down-regulate endogenous IDH1
`and determined the a-KG levels in U-87MG
`
`human glioblastoma cells. Two independent
`short hairpin RNAs (shRNAs) decreased IDH1
`mRNA by more than 75% and reduced cellular
`a-KG levels by up to 50% (fig. S3). Expression
`of the IDH1R132H mutant at a level similar to the
`endogenous protein (fig. S4A) in the cytoplasm
`of U-87MG cells caused a dose-dependent re-
`duction of a-KG levels (Fig. 2D) (fig. S4, A and
`
`these data indicate that tumor-
`B). Together,
`derived mutant IDH1 dominantly inhibits the
`wild-type IDH1 by forming a catalytically in-
`active heterodimer, resulting in a decrease of
`cellular a-KG.
`Because a-KG is required by prolylhydrox-
`ylases (PHD), enzymes that hydroxylate and
`promote the degradation of hypoxia-inducible
`
`REPORTS
`
`Fig. 2. The R132H mutation dominantly inhibits
`IDH1 activity and reduces cellular levels of a-KG.
`(A) The WT:R132H heterodimer of IDH1 has low
`specific activity. The specific activities of WT:WT,
`R132H:R132H, and WT:R132H dimers were mea-
`sured under conditions of NADP+ (10 mM), ICT (30 mM),
`and 2 mM MnCl2. Activities were normalized by pro-
`tein levels, and wild-type activity was arbitrarily set as
`100%. Shown are mean values of triplicate experi-
`ments TSD. (B) A close-up view showing the con-
`formational differences between the IDH1-NADP+
`and IDH1-NADP+-ICT complexes at the active site.
`The enzyme adopts a quasi-open conformation in the
`IDH1-NADP+ complex (cyan) and a closed conforma-
`tion in the IDH1-NADP+-ICT complex (yellow). The
`bound ICT and the side chains of several residues in
`IDH1 involved in ICT binding are shown. (C) The
`WT:R132H heterodimer loses cooperative binding
`to ICT. The activities of the WT:WT and WT:R132H
`enzymes were assayed with increasing concentra-
`tions of ICT in the presence of 100 mM NADP+ and
`2 mM Mn2+. Shown are mean values of duplicate
`assays TSD. The inset is an expanded view showing
`the IDH1 activities at lower isocitrate concentrations.
`(D) Cellular a-KG levels decrease with increasing
`IDH1R132H expression. The upper panel is a Western
`blot showing expression levels of the transfected
`IDH1R132H mutant in U-87MG cells. The a-KG level
`in cells transfected with empty vector was set as
`100%, and this value was used to calculate the
`relative a-KG level in cells transfected with differ-
`ent amounts of IDH1R132H mutant. Shown are mean
`values of triplicate assays TSD.
`
`Fig. 3. a-KG mediates the HIF-1a induction in cells with a decreased
`IDH1 activity. (A) IDH1 knockdown elevates HIF-1a levels in U-87MG
`glioblastoma cells. IDH1 and HIF-1a protein levels were determined by
`Western blotting from stable U-87MG cells transduced with empty
`retrovirus or retrovirus expressing different shRNAs silencing IDH1. (B)
`Ectopic expression of the IDH1R132H mutant elevates HIF-1a levels in U-
`87MG and HEK293T cells. The IDH1R132H mutant was overexpressed in
`U-87MG or HEK293T cells, and protein levels were detected by Western
`blot. CoCl2-treated cells (a mimetic of hypoxia) and cells overexpressing
`wild-type IDH1 were also included as controls. (C) A cell-permeable a-KG
`derivative blocks HIF-1a induction in cells expressing IDH1R132H. The U-
`87MG cells were transfected with IDH1R132H, and different concen-
`trations of octyl-a-KG ester were added to each transfected cell for 4
`hours. HIF-1a protein levels were assayed by Western blot.
`
`www.sciencemag.org SCIENCE VOL 324
`
`10 APRIL 2009
`
`263
`
`Rigel Exhibit 1012
`Page 3 of 5
`
`

`

`REPORTS
`
`statistically stronger HIF-1a signal than did 12
`tumors that did not harbor this mutation (Fig.
`4C). In IDH1 mutated tumors, 28.1 T 6.7% of
`the cells stained positive for HIF-1a, whereas
`in tumors with wild-type IDH1, only 15.8 T
`3.5% of the cells stained positive (P < 0.001)
`(Fig. 4C). IDH1-mutated gliomas also exhibited
`an increase in VEGF levels compared with gliomas
`without IDH1 mutation of similar type and grade
`(fig. S7B).
`In summary, we have shown that IDH1 is
`likely to function as a tumor suppressor gene
`rather than as an oncogene. The glioma-associated
`mutations dominantly inhibit the activity of wild-
`type IDH1 through heterodimer formation (fig. S8).
`IDH1 gene mutations in gliomas exhibit
`two
`unique features: the lack of LOH and the lack
`of apparent inactivating mutations such as frame-
`shift or truncations. Our findings help to explain
`both features, as dominant
`inhibition would
`eliminate the selection pressure to mutate or lose
`the remaining wild-type allele (LOH) and frame-
`shift or premature termination would likely gen-
`erate IDH1 fragments unable to dimerize with
`and inhibit the wild-type IDH1 protein.
`The link between IDH1 and HIF-1a high-
`lights an emerging theme in which mutationally
`altered metabolic enzymes are thought to con-
`tribute to tumor growth by stimulating the HIF-1a
`pathway and tumor angiogenesis. The genes en-
`coding two TCA enzymes, fumarate hydratase
`(FH) and succinate dehydrogenase (SDH), have
`been found to sustain loss-of-function mutations
`in certain human tumors, which likewise corre-
`late with an increase in HIF-1a levels (10, 11). In
`addition to affecting PHD, an alteration in a-KG
`might contribute to tumorigenesis by affecting
`other dioxygenases that use a-KG as a substrate.
`IDH1 also catalyzes the production of NADPH;
`thus, it is possible that a reduction in NADPH
`levels resulting from IDH1 mutation contributes
`to tumorigenesis through effects on cell metabo-
`lism and growth. Several dozen anticancer agents
`directly targeting HIF-1a are under development
`or being tested (12). Our finding that an a-KG
`derivative can reverse the induction of HIF-1a
`levels in cultured cells expressing mutant IDH1
`suggests that drugs mimicking a-KG may merit
`exploration as a therapy for gliomas that harbor
`an IDH1 mutation.
`
`References and Notes
`1. F. B. Furnari et al., Genes Dev. 21, 2683 (2007).
`2. D. W. Parsons et al., Science 321, 1807 (2008).
`3. J. Balss et al., Acta Neuropathol. 116, 597 (2008).
`4. F. E. Bleeker et al., Hum. Mutat. 30, 7 (2009).
`5. H. Yan et al., N. Engl. J. Med. 360, 765 (2009).
`6. B. S. Winkler, N. DeSantis, F. Solomon, Exp. Eye Res. 43,
`829 (1986).
`7. X. Xu et al., J. Biol. Chem. 279, 33946 (2004).
`8. S. Soundar, B. L. Danek, R. F. Colman, J. Biol. Chem.
`275, 5606 (2000).
`9. K. R. Albe, M. H. Butler, B. E. Wright, J. Theor. Biol. 143,
`163 (1990).
`10. E. D. MacKenzie et al., Mol. Cell. Biol. 27, 3282
`(2007).
`11. O. C. Ingebretsen, Biochim. Biophys. Acta 452, 302
`(1976).
`
`Fig. 4. IDH1 activity affects the levels of HIF-1a and HIF-1a target genes in gliomas and cultured cells.
`(A) Overexpression of the IDH1R132H mutant in U-87MG cells stimulates expression of HIF-1a target genes
`(Glut1, VEGF, and PGK1) as assayed by QPCR. Shown are mean values of triplicate assays TSD. (B)
`Inhibition of IDH1 by oxalomalate activates HIF-1a target genes. U-87MG cells were either untreated
`(control), treated with 5 mM oxalomalate, an IDH1 inhibitor, or treated with CoCl2, a hypoxia mimetic.
`HIF-1a target gene mRNAs were determined. Shown are mean values of triplicate assays TSD. (C)
`Immunohistochemistry of HIF-1a was carried out in 12 human gliomas with wild-type IDH1 and 8 gliomas
`of similar grade harboring a mutated IDH1 allele. Shown are side-by-side comparisons of four gliomas
`representing different types or grades. Scale bar, 40 mM. Five fields (~173 mm2 each) were randomly
`selected from each sample for quantification of HIF-1a-positive staining area. Statistical analysis was
`performed using seven IDH1 wild-type and seven IDH1-mutated gliomas.
`
`factor 1a (HIF-1a), we hypothesized that de-
`creased IDH1 activity might stabilize HIF-1a and
`increase its steady-state levels. We found that
`HIF-1a protein levels in U-87MG cells were
`elevated in response to shRNA-mediated knock-
`down of IDH1 (Fig. 3A). Conversely, overexpres-
`sion of wild-type IDH1 reduced HIF-1a protein
`levels in HeLa (fig. S5A) and U-87MG cells
`(Fig. 3B). Notably, overexpression of IDH1R132H
`mutant increased HIF-1a protein levels in U-
`87MG and HEK293T cells (Fig. 3B). These
`results suggest that IDH1 regulates HIF-1a levels
`by controlling the level of a-KG. We tested this
`hypothesis by treating cells with octyl-a-KG, a
`cell-permeable derivative of a-KG that upon
`entering the cells is converted into a-KG after
`hydrolysis of the ester group (10). We found that
`octyl-a-KG suppressed the HIF-1a induction
`caused by either IDH1 knockdown in HeLa cells
`(fig. S5B) or overexpression of IDH1R132H mu-
`tant in U-87MG cells (Fig. 3C). We therefore con-
`clude that a reduction in IDH1 activity produces a
`reduction in a-KG levels that in turn can lead to
`stabilization of HIF-1a.
`We next determined whether inhibition of
`IDH1 enzyme activity leads to up-regulated ex-
`pression of HIF-1a target genes. HIF-1a is a key
`
`component of HIF-1, a transcription factor that
`senses low cellular oxygen levels and that regu-
`lates the expression of genes implicated in glucose
`metabolism, angiogenesis, and other signaling
`pathways that are critical to tumor growth. Quan-
`titative real-time fluorescence polymerase chain
`reaction (QPCR) of mRNAs corresponding to
`three well-established HIF-1a target genes, glu-
`cose transporter 1 (Glut1), vascular endothelial
`growth factor (VEGF), and phosphoglycerate
`kinase (PGK1) showed that IDH1 knockdown
`induced the expression of these HIF-1a target
`genes (fig. S6A). Moreover, expression of the
`IDH1R132H mutant, but not wild-type IDH1,
`strongly induced HIF-1a target gene expression
`(Fig. 4A). Oxalomalate, a competitive inhibitor
`of IDH1 (11) (fig. S6B), also induced expression
`of these HIF-1a target genes (Fig. 4B).
`Finally, we determined whether IDH1 muta-
`tions correlate with elevated levels of HIF-1a in
`human gliomas. In a collection of 26 glioma
`samples, we identified 8 tumors that contained
`the R132H mutation in one allele of IDH1 (table
`S1) (fig. S7A). Using immunohistochemistry, we
`compared HIF-1a expression in gliomas with
`and without IDH1 mutations. We found that 8
`tumors harboring the R132H mutation showed a
`
`264
`
`10 APRIL 2009 VOL 324 SCIENCE www.sciencemag.org
`
`Rigel Exhibit 1012
`Page 4 of 5
`
`

`

`12. G. L. Semenza, Drug Discov. Today 12, 853 (2007).
`13. We thank members of the Fudan Molecular and Cell
`Biology Laboratory for valuable input; Y. Liu, X. Liu, and
`H. Zhu for assistance with histology; Z. Bao, L. Yang,
`Q. Shi, and G. Zhao for clinical samples; and S. Jackson
`for reading the manuscript. This work is supported by the
`985 program from the Chinese Ministry of Education,
`State Key Development Programs of China
`
`(2009CB918401, 2006CB806700), National 863
`Program of China (2006AA02A308), China NSF grants
`(30600112 and 30871255) and Shanghai Key Basic
`Research Projects (06JC14086, 07PJ14011, and
`08JC1400900), and NIH grants (to K.-L.G. and Y.X.).
`Y. Xiong, K.-L. Guan, and S. Zhao are applying for
`a patent related to the work on permeable
`alpha-ketogluterate.
`
`Supporting Online Material
`www.sciencemag.org/cgi/content/full/324/5924/261/DC1
`Materials and Methods
`Figs. S1 to S8
`Table S1
`
`15 January 2009; accepted 10 March 2009
`10.1126/science.1170944
`
`REPORTS
`
`Demonstration of Genetic Exchange
`During Cyclical Development of
`Leishmania in the Sand Fly Vector
`Natalia S. Akopyants,1* Nicola Kimblin,2* Nagila Secundino,2 Rachel Patrick,2 Nathan Peters,2
`Phillip Lawyer,2 Deborah E. Dobson,1 Stephen M. Beverley,1† David L. Sacks2†‡
`
`Genetic exchange has not been shown to be a mechanism underlying the extensive diversity of
`Leishmania parasites. We report here evidence that the invertebrate stages of Leishmania are
`capable of having a sexual cycle consistent with a meiotic process like that described for African
`trypanosomes. Hybrid progeny were generated that bore full genomic complements from both
`parents, but kinetoplast DNA maxicircles from one parent. Mating occurred only in the sand fly
`vector, and hybrids were transmitted to the mammalian host by sand fly bite. Genetic exchange
`likely contributes to phenotypic diversity in natural populations, and analysis of hybrid progeny
`will be useful for positional cloning of the genes controlling traits such as virulence, tissue
`tropism, and drug resistance.
`
`Parasitic protozoa of the genus Leishmania
`
`cause a spectrum of human diseases that
`pose serious public health challenges for
`prevention, diagnosis, and treatment. The diver-
`sity of Leishmania species, with more than 20
`currently recognized, is thought to have arisen by
`gradual accumulation of divergent mutations rather
`than by sexual recombination. Tibayrenc et al.
`(1) have reported strong linkage disequilibrium
`in several Leishmania species and proposed that
`these parasites are essentially clonal. This notion
`must be reconciled, however, with the accumu-
`lating examples of naturally occurring strains that
`share genotypic markers from two recognized
`species and thereby provide circumstantial evi-
`dence for sexual recombination (2–4). Genetic
`exchange has been documented for the other
`trypanosomatids that cause human disease.
`Hybrid genotypes were observed in tsetse flies
`during cotransmission of two strains of Trypano-
`soma brucei (5) and in mammalian cells after
`coinfection with two clones of Trypanosoma
`cruzi differing in drug-resistance markers (6).
`Using drug resistance markers, we provide evi-
`dence for genetic exchange in Leishmania
`major and discuss the implications of these
`findings to Leishmania biology and experimen-
`tal analysis.
`
`1Department of Molecular Microbiology, Washington Univer-
`sity School of Medicine, St. Louis, MO 63110, USA. 2Labora-
`tory of Parasitic Diseases, National Institute of Allergy and
`Infectious Diseases, NIH, Bethesda, MD 20892, USA.
`*†These authors contributed equally to this work.
`‡To whom correspondence should be addressed. E-mail:
`dsacks@nih.gov
`
`One parental clone, LV39c5(HYG), was de-
`rived from strain LV39 clone 5 (MHOM/SU/59/P)
`and was heterozygous for an allelic replacement of
`the LPG5A on chromosome 24 by a hygromycin
`B–resistance cassette (LPG5A/LPG5A::ΔHYG)
`(7). The second parental clone, FV1(SAT), was
`derived from NIH Friedlin clone V1 (MHOM/
`IL/80/FN) and bore a heterozygous nourseothricin–
`resistance (SAT) marker, integrated along with a
`linked firefly luciferase (LUC) reporter gene into
`one allele of the ~24 rRNA cistrons located on
`chromosome 27 (8) (+/SSU::SAT-LUC). These
`strains were chosen as they are phenotypically
`identical to their respective parental wild-type (WT)
`virulent L. major; whereas the markers were chosen
`because they are functionally independent (9). The
`target gene modifications were chosen because
`they caused no effect on normal growth in vitro
`or in mouse infections (10), and epistatic interac-
`tions were not anticipated between these alleles.
`Multiple attempts to generate hybrid para-
`sites resistant to both antibiotics during in vitro co-
`culture of the parental lines were unsuccessful (11).
`The parental clones were tested for their ability to
`generate parasites resistant to both drugs during
`coinfection in the sand fly. The growth of each
`parental line in Phlebotomus duboscqi, a natural
`vector of L. major, is shown in fig. S1. Promas-
`tigotes of each parent survived the initial period
`of blood-meal digestion and excretion (days
`1 to 6) and underwent metacyclogenesis at a
`comparable frequency (20 to 60%), although the
`FV1(SAT) parent established and maintained a
`higher intensity of infection by a factor of 3 to 4.
`The parental clones were tested for their ability
`
`to generate doubly drug–resistant parasites dur-
`ing coinfection in the sand fly. Flies were fed
`through a membrane on mouse blood contain-
`ing 3 and 1 × 106/ml of the LV39c5(HYG) and
`FV1(SAT) lines, respectively, each obtained from
`log-phase cultures and extensively washed to
`remove antibiotics. A total of 102 flies from four
`independent coinfection experiments were dis-
`sected 13 to 16 days postinfection; at this time,
`they harbored mature infections with an average
`of 39,400 T 14,700 promastigotes per midgut.
`Flies cannot be maintained under aseptic con-
`ditions, and more than half of the cultures es-
`tablished from the midgut parasites were lost to
`fungal contamination during the subsequent 1 to
`2 weeks of culture. In the remaining cultures, 12
`(26%) grew out promastigotes that were resistant
`to both drugs. Clonal lines were generated from
`nine of the doubly drug–resistant populations,
`and the genotypes and phenotypes of one or two
`clones from each culture were determined (sum-
`marized in Table 1).
`Polymerase chain reaction (PCR) tests with
`primers specific for the parental markers showed
`that all doubly drug–resistant clones tested con-
`tained both the HYG and SAT drug-resistance
`genes (Fig. 1A, Table 1, and table S3). Controls
`showed that the marker loci had not rearranged
`dur

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